thanks: These authors contributed equally to this workthanks: These authors contributed equally to this work

Observation of Momentum Space Josephson Effects

Annesh Mukhopadhyay Department of Physics and Astronomy, Washington State University, Pullman, WA 99164-2814, USA    Xi-Wang Luo CAS Key Laboratory of Quantum Information, University of Science and Technology of China, Hefei, Anhui 230026, China Synergetic Innovation Center of Quantum Information and Quantum Physics, University of Science and Technology of China, Hefei, Anhui 230026, China Hefei National Laboratory, University of Science and Technology of China, Hefei 230088, China    Colby Schimelfenig Department of Physics and Astronomy, Washington State University, Pullman, WA 99164-2814, USA    M. K. H. Ome Department of Physics and Astronomy, Washington State University, Pullman, WA 99164-2814, USA    Sean Mossman Department of Physics and Astronomy, Washington State University, Pullman, WA 99164-2814, USA Department of Physics and Biophysics, University of San Diego, San Diego, CA 92110, USA    Chuanwei Zhang [email protected] Department of Physics, The University of Texas at Dallas, Richardson, TX 75080-3021, USA Department of Physics, Washington University in St. Louis, St. Louis, MO 63130, USA    Peter Engels [email protected] Department of Physics and Astronomy, Washington State University, Pullman, WA 99164-2814, USA
Abstract

The momentum space Josephson effect describes the supercurrent flow between weakly coupled Bose-Einstein condensates (BECs) at two discrete momentum states. Here, we experimentally observe this exotic phenomenon using a BEC with Raman-induced spin-orbit coupling, where the tunneling between two local band minima is implemented by the momentum kick of an additional optical lattice. A sudden quench of the Raman detuning induces coherent spin-momentum oscillations of the BEC, which is analogous to the a.c. Josephson effect. We observe both plasma and regular Josephson oscillations in different parameter regimes. The experimental results agree well with the theoretical model and numerical simulation, and showcase the important role of nonlinear interactions. We also show that the measurement of the Josephson plasma frequency gives the Bogoliubov zero quasimomentum gap, which determines the mass of the corresponding pseudo-Goldstone mode, a long-sought phenomenon in particle physics. The observation of momentum space Josephson physics offers an exciting platform for quantum simulation and sensing utilizing momentum states as a synthetic degree.

Introduction.

The Josephson effect describes supercurrents flowing between two reservoirs with a weak tunneling link (e.g., flow through a thin insulating barrier) [1, 2]. Josephson effects have experimentally been observed in many platforms, ranging from solid state superconductors [3] to superfluid Helium [4, 5, 6, 7, 8], exciton polaritons [9], and ultra-cold atomic gases [10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 23, 24, 25, 26]. Important applications of Josephson effects include superconducting quantum interference devices (SQUIDs) [27, 28], superconducting qubits [29, 30, 31, 32], and precision measurements [27].

In recent years, momentum states of ultracold bosons have emerged as a new synthetic degree of freedom for quantum matter and simulation. In this context, the Josephson effect in momentum space has been theoretically predicted for BECs located at two momentum states with a weak coupling induced by momentum kicks of laser beams [33]. Such momentum space tunneling has been implemented in experiments using a Bragg transition for a single component BEC [34, 35] or an optical lattice in a spin-orbit coupled BEC [36]. Despite significant experimental progress in the observation of various forms of quantum dynamics in momentum space lattices (e.g., macroscopic quantum self-trap** or phase-driven nonlinear dynamics [34, 35]), the momentum space Josephson oscillation has not been observed in experiments due to the challenge of realizing a coherent ground state BEC occupying two momentum states with a long lifetime.

In this Letter, we show experimental evidence for the momentum space Josephson effect in a spin-orbit coupled BEC [37, 38, 39], whose double well band dispersion possesses two band minima at different momentum states, in analogy to real space Josephson junctions. The incorporation of a weak optical lattice induces a coupling between BECs located at two band minima, leading to the experimental observation of the long-lived (>100absent100>100> 100 ms) superfluid stripe ground state [36]. Starting from the stripe ground state, a supercurrent through the momentum space junction is induced by a sudden quench of the Raman detuning between two band minima, similar to applying a voltage in a superconducting Josephson junction. The detuning quench displaces the initial stripe state from the ground state for the final detuning parameter, leading to periodic spin-momentum oscillations observed in experiments that are Josephson oscillations.

We observe two types of Josephson oscillations: i) Josephson plasma oscillations, which are characterized by a small change in population and small phase differences between the two BECs, excited through a weak change of the system ground state; ii) regular Josephson oscillations with a large population oscillation and a continuous increase (or decrease) of the phase difference, excited through a large change of the ground state. Our experimental results show good agreement with a theoretical model based on a two-mode approximation and numerical simulation based on the nonlinear Gross-Pitaevskii (GP) equation. The observed constant plateau of the plasma oscillation frequency in the weak lattice region showcases the important role of nonlinear interactions. Furthermore, we find that the observed Josephson plasma oscillation frequency corresponds to the zero quasimomentum gap of the Bogoliubov excitations in the superfluid stripe phase, which, in our system, represents the mass of a pseudo-Goldstone mode [40] emerging from explicit symmetry breaking (here, the weak optical lattice breaks spatial translational symmetry). Pseudo-Goldstone modes, first proposed in particle physics [41], have been a long-sought phenomenon in many different fields, and our work provides one direct experimental evidence for observing such an exotic mode.

Description of the system.

The experimental setup for the spin-orbit coupled BEC [[SeeSupplementaryMaterialsat][formoretheoreticaldetailsaboutthemodel, Josephson&GPdynamics, therelationwithBogoliubovspectrum, aswellassomeexperimentaldetailsabouttheparametersandBraggspectrum.]sup_mat] has been described in our previous work [36]. Briefly, a 87Rb BEC is confined in a cigar-shaped crossed optical dipole trap (Fig. 1 (a)). An external magnetic field applied along the x𝑥xitalic_x-axis lifts the degeneracy among the three Zeeman states (mFsubscript𝑚𝐹m_{F}italic_m start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT) in the F=1𝐹1F=1italic_F = 1 hyperfine manifold. A pair of \qty789nm Raman beams intersecting at approximately 45°45°45 ⁢ ° angles with the x𝑥xitalic_x-axis couples the ||1,1\lvert\uparrow\rangle\equiv|1,-1\rangle| ↑ ⟩ ≡ | 1 , - 1 ⟩ and ||1,0\lvert\downarrow\rangle\equiv|1,0\rangle| ↓ ⟩ ≡ | 1 , 0 ⟩ Zeeman split states (Fig. 1 (b)), which generates spin-orbit coupling (SOC) in the x𝑥xitalic_x-direction. The Raman coupling provides an effective momentum offset of 2kR2Planck-constant-over-2-pisubscript𝑘R2\hbar k_{\text{R}}2 roman_ℏ italic_k start_POSTSUBSCRIPT R end_POSTSUBSCRIPT between these two pseudospin states. Due to the quadratic Zeeman splitting, the |1,+1ket11|1,+1\rangle| 1 , + 1 ⟩ state is sufficiently decoupled and does not play a significant role [[SeeSupplementaryMaterialsat][formoretheoreticaldetailsaboutthemodel, Josephson&GPdynamics, therelationwithBogoliubovspectrum, aswellassomeexperimentaldetailsabouttheparametersandBraggspectrum.]sup_mat]. Additionally, two \qty1064nm laser beams co-propagating with the Raman beams create a weak stationary optical lattice VL(x)=2ΩLsin2(kLx)subscript𝑉L𝑥2Planck-constant-over-2-pisubscriptΩLsuperscript2subscript𝑘L𝑥V_{\text{L}}(x)=2\hbar\Omega_{\text{L}}\sin^{2}(k_{\text{L}}x)italic_V start_POSTSUBSCRIPT L end_POSTSUBSCRIPT ( italic_x ) = 2 roman_ℏ roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT roman_sin start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_k start_POSTSUBSCRIPT L end_POSTSUBSCRIPT italic_x ) along the x𝑥xitalic_x-direction, which provides a 2kL2Planck-constant-over-2-pisubscript𝑘L2\hbar k_{\text{L}}2 roman_ℏ italic_k start_POSTSUBSCRIPT L end_POSTSUBSCRIPT momentum kick while kee** the spin unchanged. Additional experimental details involving the atomic states and energy scales are provided in [[SeeSupplementaryMaterialsat][formoretheoreticaldetailsaboutthemodel, Josephson&GPdynamics, therelationwithBogoliubovspectrum, aswellassomeexperimentaldetailsabouttheparametersandBraggspectrum.]sup_mat].

Refer to caption
Figure 1: Illustration of the experimental setup and the momentum space Josephson effect. (a) A crossed optical dipole trap (red) with two Raman laser beams (green) co-linear with two optical lattice beams (red) intersecting at the BEC position in the center. (b) Two-photon Raman transitions within the F=1𝐹1F=1italic_F = 1 hyperfine manifold of 87Rb. (c) Band structure of HSOCsubscript𝐻SOCH_{\text{SOC}}italic_H start_POSTSUBSCRIPT SOC end_POSTSUBSCRIPT for ΩR=2.7ERPlanck-constant-over-2-pisubscriptΩR2.7subscript𝐸R\hbar\Omega_{\text{R}}=2.7\,E_{\text{R}}roman_ℏ roman_Ω start_POSTSUBSCRIPT R end_POSTSUBSCRIPT = 2.7 italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT and δ=2π×\qty500Hz𝛿2𝜋\qty500𝐻𝑧\delta=2\pi\times\qty{500}{Hz}italic_δ = 2 italic_π × 500 italic_H italic_z. (d) Phase space diagram demonstrating Josephson dynamics for ΩL=0.5ERPlanck-constant-over-2-pisubscriptΩL0.5subscript𝐸R\hbar\Omega_{\text{L}}=0.5\,E_{\text{R}}roman_ℏ roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT = 0.5 italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT, and gn=0.25ER𝑔𝑛0.25subscript𝐸Rgn=0.25\,E_{\text{R}}italic_g italic_n = 0.25 italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT.

The dynamics of the system can be described by the one-dimensional GP equation

itψ=[H0+g|ψ(x,t)|2]ψ,𝑖subscript𝑡𝜓delimited-[]subscript𝐻0𝑔superscript𝜓𝑥𝑡2𝜓i\partial_{t}\psi=\left[H_{0}+g|\psi(x,t)|^{2}\right]\psi,italic_i ∂ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_ψ = [ italic_H start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_g | italic_ψ ( italic_x , italic_t ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] italic_ψ , (1)

where ψ=(ψ,ψ)T𝜓superscriptsubscript𝜓subscript𝜓𝑇\psi=\left(\psi_{\uparrow},\psi_{\downarrow}\right)^{T}italic_ψ = ( italic_ψ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT , italic_ψ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT is the two-component spinor wavefunction with the normalization to the total number of atoms N=|ψ|2𝑑x𝑁superscript𝜓2differential-d𝑥N=\int\left|\psi\right|^{2}dxitalic_N = ∫ | italic_ψ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_d italic_x, g𝑔gitalic_g is the density interaction strength, and H0=HSOC+VL(x)subscript𝐻0subscript𝐻SOCsubscript𝑉L𝑥H_{0}=H_{\text{SOC}}+{V}_{\text{L}}(x)italic_H start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = italic_H start_POSTSUBSCRIPT SOC end_POSTSUBSCRIPT + italic_V start_POSTSUBSCRIPT L end_POSTSUBSCRIPT ( italic_x ) is the single particle Hamiltonian with

HSOC=(ix+σz)2δ2σz+ΩR2σx.subscript𝐻SOCsuperscript𝑖subscript𝑥subscript𝜎𝑧2𝛿2subscript𝜎𝑧subscriptΩR2subscript𝜎𝑥H_{\text{SOC}}=(i\partial_{x}+\sigma_{z})^{2}-\frac{\delta}{2}\sigma_{z}+\frac% {\Omega_{\text{R}}}{2}\sigma_{x}.italic_H start_POSTSUBSCRIPT SOC end_POSTSUBSCRIPT = ( italic_i ∂ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT + italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - divide start_ARG italic_δ end_ARG start_ARG 2 end_ARG italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT + divide start_ARG roman_Ω start_POSTSUBSCRIPT R end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG italic_σ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT . (2)

Here ΩRsubscriptΩR\Omega_{\text{R}}roman_Ω start_POSTSUBSCRIPT R end_POSTSUBSCRIPT is the Raman coupling strength, δ𝛿\deltaitalic_δ is the detuning of the two-photon Raman transition, and kRPlanck-constant-over-2-pisubscript𝑘R\hbar k_{\text{R}}roman_ℏ italic_k start_POSTSUBSCRIPT R end_POSTSUBSCRIPT and ER=2kR22m=h×\qty1.96kHzsubscript𝐸RsuperscriptPlanck-constant-over-2-pi2superscriptsubscript𝑘R22𝑚\qty1.96𝑘𝐻𝑧E_{\text{R}}=\frac{\hbar^{2}k_{\text{R}}^{2}}{2m}=h\times\qty{1.96}{kHz}italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT = divide start_ARG roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_k start_POSTSUBSCRIPT R end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_m end_ARG = italic_h × 1.96 italic_k italic_H italic_z are the momentum and energy units.

Fig. 1 (c) shows the momentum-space double-well band dispersion of HSOCsubscript𝐻SOCH_{\text{SOC}}italic_H start_POSTSUBSCRIPT SOC end_POSTSUBSCRIPT. In the experiment, the period of the optical lattice is set such that 2kL2Planck-constant-over-2-pisubscript𝑘L2\hbar k_{\text{L}}2 roman_ℏ italic_k start_POSTSUBSCRIPT L end_POSTSUBSCRIPT equals the separation between the two spin-orbit band minima [36]. The optical lattice leads to the hop** between the BECs at the two band minima, in analogy to the tunneling between two superconductors separated by an insulating barrier in a Josephson junction. While the optical lattice produces multiple off-resonance couplings in addition, the momentum space Josephson junction can be more intuitively understood using a two-mode approximation, i.e., considering only two BEC modes at two band minima with ψ=(ϕlχleikLx+ϕrχreikLx)eikbx𝜓subscriptitalic-ϕ𝑙subscript𝜒𝑙superscript𝑒𝑖subscript𝑘L𝑥subscriptitalic-ϕ𝑟subscript𝜒𝑟superscript𝑒𝑖subscript𝑘L𝑥superscript𝑒𝑖subscript𝑘b𝑥\psi=\left(\phi_{l}\chi_{l}e^{-ik_{\text{L}}x}+\phi_{r}\chi_{r}e^{ik_{\text{L}% }x}\right)e^{ik_{\text{b}}x}italic_ψ = ( italic_ϕ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT italic_χ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_i italic_k start_POSTSUBSCRIPT L end_POSTSUBSCRIPT italic_x end_POSTSUPERSCRIPT + italic_ϕ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT italic_χ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_k start_POSTSUBSCRIPT L end_POSTSUBSCRIPT italic_x end_POSTSUPERSCRIPT ) italic_e start_POSTSUPERSCRIPT italic_i italic_k start_POSTSUBSCRIPT b end_POSTSUBSCRIPT italic_x end_POSTSUPERSCRIPT, where χjsubscript𝜒𝑗\chi_{j}italic_χ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT are the spinor wavefunction at two band minima, ϕj(t)subscriptitalic-ϕ𝑗𝑡\phi_{j}(t)italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t ) are the mode population coefficients, and kbsubscript𝑘bk_{\text{b}}italic_k start_POSTSUBSCRIPT b end_POSTSUBSCRIPT is the bias momentum induced by the detuning δ𝛿\deltaitalic_δ. This model also neglects modes in the excited band. Denoting ϕj=njeiθjsubscriptitalic-ϕ𝑗subscript𝑛𝑗superscript𝑒𝑖subscript𝜃𝑗\phi_{j}=\sqrt{n_{j}}e^{i\theta_{j}}italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = square-root start_ARG italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG italic_e start_POSTSUPERSCRIPT italic_i italic_θ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUPERSCRIPT, the GP equation can be projected as

τzsubscript𝜏𝑧\displaystyle\partial_{\tau}{z}∂ start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT italic_z =\displaystyle== 1z2sinθ1superscript𝑧2𝜃\displaystyle-\sqrt{1-z^{2}}\sin{\theta}- square-root start_ARG 1 - italic_z start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG roman_sin italic_θ
τθsubscript𝜏𝜃\displaystyle\partial_{\tau}{\theta}∂ start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT italic_θ =\displaystyle== Λz+z1z2cosθΔEΛ𝑧𝑧1superscript𝑧2𝜃Δ𝐸\displaystyle\Lambda z+\frac{z}{\sqrt{1-z^{2}}}\cos{\theta}-\Delta Eroman_Λ italic_z + divide start_ARG italic_z end_ARG start_ARG square-root start_ARG 1 - italic_z start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_ARG roman_cos italic_θ - roman_Δ italic_E (3)

in terms of the phase difference θ=θlθr𝜃subscript𝜃𝑙subscript𝜃𝑟\theta=\theta_{l}-\theta_{r}italic_θ = italic_θ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT - italic_θ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT and relative population difference z=(nrnl)/n𝑧subscript𝑛𝑟subscript𝑛𝑙𝑛z=\left(n_{r}-n_{l}\right)/nitalic_z = ( italic_n start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT - italic_n start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ) / italic_n, where n=nl+nr𝑛subscript𝑛𝑙subscript𝑛𝑟n=n_{l}+n_{r}italic_n = italic_n start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT + italic_n start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT is taken as a constant by neglecting populations in other modes. τ=2Kt𝜏2𝐾𝑡\tau=2Ktitalic_τ = 2 italic_K italic_t, where K=ΩL2χlχr𝐾subscriptΩL2superscriptsubscript𝜒𝑙subscript𝜒𝑟K=\frac{\Omega_{\text{L}}}{2}\chi_{l}^{\ast}\chi_{r}italic_K = divide start_ARG roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG italic_χ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT italic_χ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT describes the hop** between the two modes (K𝐾Kitalic_K is chosen to be real without loss of generality). Λ=Un/(2K)Λ𝑈𝑛2𝐾\Lambda=-Un/(2K)roman_Λ = - italic_U italic_n / ( 2 italic_K ), where U=g|χlχr|2𝑈𝑔superscriptsuperscriptsubscript𝜒𝑙subscript𝜒𝑟2U=g|\chi_{l}^{\ast}\chi_{r}|^{2}italic_U = italic_g | italic_χ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT italic_χ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT represents the interaction strength of the BECs with two modes. ΔE=(ElEr)/(2K)Δ𝐸subscript𝐸𝑙subscript𝐸𝑟2𝐾\Delta E=(E_{l}-E_{r})/(2K)roman_Δ italic_E = ( italic_E start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ) / ( 2 italic_K ), with Ej=𝑑xχjH0χj+(g+U)nsubscript𝐸𝑗differential-d𝑥superscriptsubscript𝜒𝑗subscript𝐻0subscript𝜒𝑗𝑔𝑈𝑛E_{j}=\int dx\chi_{j}^{\ast}H_{0}\chi_{j}+(g+U)nitalic_E start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = ∫ italic_d italic_x italic_χ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT italic_H start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_χ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT + ( italic_g + italic_U ) italic_n, is the energy difference between the two modes. Eq. (3) describes a Bose Josephson junction governed by the effective Hamiltonian Heff=Λ2z21z2cosθΔEzsubscript𝐻effΛ2superscript𝑧21superscript𝑧2𝜃Δ𝐸𝑧H_{\text{eff}}=\frac{\Lambda}{2}z^{2}-\sqrt{1-z^{2}}\cos{\theta}-\Delta Ezitalic_H start_POSTSUBSCRIPT eff end_POSTSUBSCRIPT = divide start_ARG roman_Λ end_ARG start_ARG 2 end_ARG italic_z start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - square-root start_ARG 1 - italic_z start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG roman_cos italic_θ - roman_Δ italic_E italic_z [15].

A typical phase space diagram is shown in Fig. 1 (d) to illustrate the momentum space Josephson dynamics. The red fixed point (z0,θ0subscript𝑧0subscript𝜃0z_{0},\theta_{0}italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT) corresponds to the equilibrium ground state that can be obtained by finding the minima of Heffsubscript𝐻effH_{\text{eff}}italic_H start_POSTSUBSCRIPT eff end_POSTSUBSCRIPT. When the BEC is initially prepared away from (z0,θ0subscript𝑧0subscript𝜃0z_{0},\theta_{0}italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT), (z,θ𝑧𝜃z,\thetaitalic_z , italic_θ) oscillate following periodic orbits in phase space, corresponding to Josephson oscillation between BECs at two band minima. There are two different types of oscillating behavior: i) when the BEC is initially prepared not far from the fixed point, θ𝜃{\theta}italic_θ only changes within a small range for the closed periodic orbits, representing Josephson plasma oscillation shown as the orange dashed line; ii) when the BEC is far from the fixed point, θ𝜃{\theta}italic_θ increases (or decreases) continuously through [0,2π]02𝜋[0,2\pi][ 0 , 2 italic_π ], corresponding to regular Josephson oscillation shown as the orange solid line [15].

Observation of Josephson oscillation.

In our experiments, we observe momentum space Josephson dynamics after a sudden quench of the Raman detuning from an initial value δisubscript𝛿𝑖\delta_{i}italic_δ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT to a final value δfsubscript𝛿𝑓\delta_{f}italic_δ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT, in analogy to the voltage-driven ac Josephson effect. After the quench, the initially prepared superstripe state at δisubscript𝛿𝑖\delta_{i}italic_δ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is no longer the ground state at δfsubscript𝛿𝑓\delta_{f}italic_δ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT. The BEC will then evolve under the two-mode approximation along a periodic orbit around the fixed point for δfsubscript𝛿𝑓\delta_{f}italic_δ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT, demonstrating the Josephson oscillation. The starting point of the experiment is the preparation of a superfluid stripe state through Raman and optical lattice dressing of the BEC [36]. SOC is generated by adiabatically ram** on the Raman beams such that the Raman coupling strength ΩRPlanck-constant-over-2-pisubscriptΩR\hbar\Omega_{\text{R}}roman_ℏ roman_Ω start_POSTSUBSCRIPT R end_POSTSUBSCRIPT increases from 00 to 2.7ER2.7subscript𝐸R2.7\,E_{\text{R}}2.7 italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT in \qty50ms. During this time, the Raman coupling is far detuned (typically, δ=2π×(\qty5.5kHz±\qty100Hz\delta=2\pi\times(\qty{5.5}{kHz}\pm\qty{100}{Hz}italic_δ = 2 italic_π × ( 5.5 italic_k italic_H italic_z ± 100 italic_H italic_z)) from the resonance. The optical lattice beams are then adiabatically applied, increasing ΩLPlanck-constant-over-2-pisubscriptΩL\hbar\Omega_{\text{L}}roman_ℏ roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT from 00 to the desired strength in \qty50ms. Following that, δ𝛿\deltaitalic_δ is linearly decreased to a desired value δisubscript𝛿𝑖\delta_{i}italic_δ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT in \qty50ms. This adiabatic process prepares the BEC in the superfluid stripe ground state for δisubscript𝛿𝑖\delta_{i}italic_δ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT.

Refer to caption
Figure 2: Spin-momentum oscillation in a double-well SOC BEC. (a)–(c) Absorption images for ΩL=0.4ERPlanck-constant-over-2-pisubscriptΩL0.4subscript𝐸R\hbar\Omega_{\text{L}}=0.4\,E_{\text{R}}roman_ℏ roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT = 0.4 italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT after an evolution time of t=0.15𝑡0.15t=0.15italic_t = 0.15, 1.051.051.051.05, and \qty1.65ms\qty1.65𝑚𝑠\qty{1.65}{ms}1.65 italic_m italic_s.

In the case of a Josephson plasma oscillation, (z,θ)𝑧𝜃\left(z,\theta\right)( italic_z , italic_θ ) oscillate along a small closed orbit around (z0,θ0)subscript𝑧0subscript𝜃0\left(z_{0},\theta_{0}\right)( italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) in phase space. We choose a fixed δf=2π×\qty500Hzsubscript𝛿𝑓2𝜋\qty500𝐻𝑧\delta_{f}=2\pi\times\qty{500}{Hz}italic_δ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT = 2 italic_π × 500 italic_H italic_z and different δi=2π×(500+Q)subscript𝛿𝑖2𝜋500𝑄\delta_{i}=2\pi\times\left(500+Q\right)italic_δ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 2 italic_π × ( 500 + italic_Q ) Hz for different ΩL{0.2,0.4,,1.4}ERPlanck-constant-over-2-pisubscriptΩL0.20.41.4subscript𝐸R\hbar\Omega_{\text{L}}\in\{0.2,0.4,\cdots,1.4\}\,E_{\text{R}}roman_ℏ roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT ∈ { 0.2 , 0.4 , ⋯ , 1.4 } italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT. Proper reasoning for choosing a finite value of the final Raman detuning (δfsubscript𝛿𝑓\delta_{f}italic_δ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT) can be found in [[SeeSupplementaryMaterialsat][formoretheoreticaldetailsaboutthemodel, Josephson&GPdynamics, therelationwithBogoliubovspectrum, aswellassomeexperimentaldetailsabouttheparametersandBraggspectrum.]sup_mat]. Suitable values of the quench frequency Q𝑄Qitalic_Q are chosen such that spin oscillations are still observable, but the initial (zi,θi)subscript𝑧𝑖subscript𝜃𝑖\left(z_{i},\theta_{i}\right)( italic_z start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) do not deviate significantly from (z0,θ0)subscript𝑧0subscript𝜃0\left(z_{0},\theta_{0}\right)( italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ), leading to plasma oscillation. After the sudden quench δiδfsubscript𝛿𝑖subscript𝛿𝑓\delta_{i}\rightarrow\delta_{f}italic_δ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT → italic_δ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT, we let the BEC evolve for a time t𝑡titalic_t in the presence of the Raman and optical lattice couplings. Subsequently, the Raman and lattice beams are switched off; the BEC is released from the crossed optical dipole trap, and a \qty17.5ms long time of flight (TOF), along with a briefly applied Stern-Gerlach field, resolves the BEC into different bare spin-momentum eigenstates. In the absorption images of the BEC, the two spin states are separated vertically, and for each spin state, the momentum components are resolved horizontally [36] (Fig. 2). We measure the total spin polarization σz=(NN)/(N+N)delimited-⟨⟩subscript𝜎𝑧subscript𝑁subscript𝑁subscript𝑁subscript𝑁\langle\sigma_{z}\rangle=(N_{\uparrow}-N_{\downarrow})/(N_{\uparrow}+N_{% \downarrow})⟨ italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ⟩ = ( italic_N start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT - italic_N start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT ) / ( italic_N start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT + italic_N start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT ) at each time t𝑡titalic_t, where Nsubscript𝑁N_{\uparrow}italic_N start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT and Nsubscript𝑁N_{\downarrow}italic_N start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT are the total number of atoms in spin |delimited-|⟩\lvert\uparrow\rangle| ↑ ⟩ and |delimited-|⟩\lvert\downarrow\rangle| ↓ ⟩ respectively. Notice that σz=jnjnχj|σz|χjdelimited-⟨⟩subscript𝜎𝑧subscript𝑗subscript𝑛𝑗𝑛quantum-operator-productsubscript𝜒𝑗subscript𝜎𝑧subscript𝜒𝑗\langle\sigma_{z}\rangle=\sum_{j}\frac{n_{j}}{n}\langle\chi_{j}|\sigma_{z}|% \chi_{j}\rangle⟨ italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ⟩ = ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT divide start_ARG italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG start_ARG italic_n end_ARG ⟨ italic_χ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT | italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT | italic_χ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ⟩, which can be written as σz=a+bzdelimited-⟨⟩subscript𝜎𝑧𝑎𝑏𝑧\langle\sigma_{z}\rangle=a+bz⟨ italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ⟩ = italic_a + italic_b italic_z under the two-mode approximation, with 2a=χl|σz|χl+χr|σz|χr2𝑎quantum-operator-productsubscript𝜒𝑙subscript𝜎𝑧subscript𝜒𝑙quantum-operator-productsubscript𝜒𝑟subscript𝜎𝑧subscript𝜒𝑟2a=\langle\chi_{l}|\sigma_{z}|\chi_{l}\rangle+\langle\chi_{r}|\sigma_{z}|\chi_% {r}\rangle2 italic_a = ⟨ italic_χ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT | italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT | italic_χ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ⟩ + ⟨ italic_χ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT | italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT | italic_χ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ⟩, and 2b=χr|σz|χrχl|σz|χl2𝑏quantum-operator-productsubscript𝜒𝑟subscript𝜎𝑧subscript𝜒𝑟quantum-operator-productsubscript𝜒𝑙subscript𝜎𝑧subscript𝜒𝑙2b=\langle\chi_{r}|\sigma_{z}|\chi_{r}\rangle-\langle\chi_{l}|\sigma_{z}|\chi_% {l}\rangle2 italic_b = ⟨ italic_χ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT | italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT | italic_χ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ⟩ - ⟨ italic_χ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT | italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT | italic_χ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ⟩. Therefore, the spin polarization oscillates with the same frequency as the Josephson oscillation. For the parameters in Fig. 3, we have a=0.0493,b=0.732formulae-sequence𝑎0.0493𝑏0.732a=0.0493,b=0.732italic_a = 0.0493 , italic_b = 0.732 [[SeeSupplementaryMaterialsat][formoretheoreticaldetailsaboutthemodel, Josephson&GPdynamics, therelationwithBogoliubovspectrum, aswellassomeexperimentaldetailsabouttheparametersandBraggspectrum.]sup_mat].

Refer to caption
Figure 3: Josephson plasma oscillation after Raman detuning quench. (a)–(c) Oscillation of the spin polarization for three different lattice coupling strengths (see main text for parameters). The solid black (light blue) curves represent the sinusoidal fitting of the experimental data (GP simulation results), while the symbols with error bars are the experimental data points. (d) Comparison between the observed ΔΔ\Deltaroman_Δ (symbols with error bars) and predicted ΔΔ\Deltaroman_Δ obtained from analyzing the Bogoliubov spectrum (solid green), quench dynamics using the GP equation (thick solid green), perturbation analysis of the two-mode Josephson model (densely dotted magenta), and the quench dynamics of Josephson model (dash-dotted red). (e) Comparison of the Bogoliubov spectrum analysis for gn=0.25ER𝑔𝑛0.25subscript𝐸Rgn=0.25\,E_{\text{R}}italic_g italic_n = 0.25 italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT and final Raman detuning δf=2π×\qty500Hzsubscript𝛿𝑓2𝜋\qty500𝐻𝑧\delta_{f}=2\pi\times\qty{500}{Hz}italic_δ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT = 2 italic_π × 500 italic_H italic_z (green solid line), \qty400Hz\qty400𝐻𝑧\qty{400}{Hz}400 italic_H italic_z (green dashed line), and \qty300Hz\qty300𝐻𝑧\qty{300}{Hz}300 italic_H italic_z (green densely dash-dot-dotted line) with the experimental data points for δf=2π×\qty500Hzsubscript𝛿𝑓2𝜋\qty500𝐻𝑧\delta_{f}=2\pi\times\qty{500}{Hz}italic_δ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT = 2 italic_π × 500 italic_H italic_z. The solid orange line represents the calculated variation of ΔΔ\Deltaroman_Δ with ΩLPlanck-constant-over-2-pisubscriptΩL\hbar\Omega_{\text{L}}roman_ℏ roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT for δf=2π×\qty500Hzsubscript𝛿𝑓2𝜋\qty500𝐻𝑧\delta_{f}=2\pi\times\qty{500}{Hz}italic_δ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT = 2 italic_π × 500 italic_H italic_z and gn=0𝑔𝑛0gn=0italic_g italic_n = 0 (non-interacting case). The distinct markers represent the oscillation frequency experimentally obtained for the corresponding ΩLPlanck-constant-over-2-pisubscriptΩL\hbar\Omega_{\text{L}}roman_ℏ roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT from the time-dependent σzdelimited-⟨⟩subscript𝜎𝑧\langle\sigma_{z}\rangle⟨ italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ⟩ plots in (a)–(c) and similar additional plots in [[SeeSupplementaryMaterialsat][formoretheoreticaldetailsaboutthemodel, Josephson&GPdynamics, therelationwithBogoliubovspectrum, aswellassomeexperimentaldetailsabouttheparametersandBraggspectrum.]sup_mat]. The star marker at ΩL=0Planck-constant-over-2-pisubscriptΩL0\hbar\Omega_{\text{L}}=0roman_ℏ roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT = 0 in (d) and (e) are obtained from Bragg spectroscopic measurements on a SOC BEC [[SeeSupplementaryMaterialsat][formoretheoreticaldetailsaboutthemodel, Josephson&GPdynamics, therelationwithBogoliubovspectrum, aswellassomeexperimentaldetailsabouttheparametersandBraggspectrum.]sup_mat].

Fig. 3 (a)–(c) shows the oscillations of σzdelimited-⟨⟩subscript𝜎𝑧\langle\sigma_{z}\rangle⟨ italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ⟩ during the post-quench time t𝑡titalic_t, measured for three different lattice coupling strengths. The corresponding quench, optical lattice coupling strength, and the spin-polarization oscillation frequency (Q,ΩL,Δ)𝑄Planck-constant-over-2-pisubscriptΩLΔ(Q,\,\hbar\Omega_{\text{L}},\,\Delta)( italic_Q , roman_ℏ roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT , roman_Δ ) for the three cases are (a) (2π×\qty1.3kHz, 0.2ER,(0.287±0.007)ER2𝜋\qty1.3𝑘𝐻𝑧0.2subscript𝐸Rplus-or-minus0.2870.007subscript𝐸R2\pi\times\qty{1.3}{kHz},\,0.2\,E_{\text{R}},\,(0.287\pm 0.007)\,E_{\text{R}}2 italic_π × 1.3 italic_k italic_H italic_z , 0.2 italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT , ( 0.287 ± 0.007 ) italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT), (b) (2π×\qty200Hz, 1.0ER,(0.530±0.016)ER2𝜋\qty200𝐻𝑧1.0subscript𝐸Rplus-or-minus0.5300.016subscript𝐸R2\pi\times\qty{200}{Hz},\,1.0\,E_{\text{R}},\,(0.530\pm 0.016)\,E_{\text{R}}2 italic_π × 200 italic_H italic_z , 1.0 italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT , ( 0.530 ± 0.016 ) italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT), and (c) (2π×\qty400Hz, 1.4ER,(0.734±0.039)ER2𝜋\qty400𝐻𝑧1.4subscript𝐸Rplus-or-minus0.7340.039subscript𝐸R2\pi\times\qty{400}{Hz},\,1.4\,E_{\text{R}},\,(0.734\pm 0.039)\,E_{\text{R}}2 italic_π × 400 italic_H italic_z , 1.4 italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT , ( 0.734 ± 0.039 ) italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT), where the errors in oscillation frequencies represent the standard errors in ΔΔ\Deltaroman_Δ obtained from the sinusoidal fitting of the corresponding data set. The experimentally observed time evolution of σzdelimited-⟨⟩subscript𝜎𝑧\langle\sigma_{z}\rangle⟨ italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ⟩ agrees reasonably well with the numerical results from directly simulating the quench dynamics using the GP Eq. (1[[SeeSupplementaryMaterialsat][formoretheoreticaldetailsaboutthemodel, Josephson&GPdynamics, therelationwithBogoliubovspectrum, aswellassomeexperimentaldetailsabouttheparametersandBraggspectrum.]sup_mat]. In Fig. 3 (d), we show experimentally measured Josephson plasma oscillation frequencies with respect to ΩLPlanck-constant-over-2-pisubscriptΩL\hbar\Omega_{\text{L}}roman_ℏ roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT and their comparison with several theoretical models and numerical calculations for the interaction strength gn=0.25ER𝑔𝑛0.25subscript𝐸Rgn=0.25\,E_{\text{R}}italic_g italic_n = 0.25 italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT. The constant Josephson plasma frequency for small values of ΩLPlanck-constant-over-2-pisubscriptΩL\hbar\Omega_{\text{L}}roman_ℏ roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT indicates a nonlinear interaction in the system, as evidenced by the comparison to the non-interacting case (Fig. 3 (e)). Additionally, Fig. 3 (e) shows the dependence of the plasma oscillation frequencies (ΔΔ\Deltaroman_Δ) on the Raman detuning (δ𝛿\deltaitalic_δ), which is an experimental parameter [[SeeSupplementaryMaterialsat][formoretheoreticaldetailsaboutthemodel, Josephson&GPdynamics, therelationwithBogoliubovspectrum, aswellassomeexperimentaldetailsabouttheparametersandBraggspectrum.]sup_mat]. The different theoretical models and numerical calculations are described below:

Refer to caption
Figure 4: Regular Josephson oscillation after Raman detuning quench. (a) Oscillation of the spin polarization. The solid black (light blue) curve represents a sinusoidal fit of the experimental data (GP simulation results). The dark blue curve represents the GP simulation results for a Raman detuning quench of δi=2π×\qty30Hzsubscript𝛿𝑖2𝜋\qty30𝐻𝑧\delta_{i}=2\pi\times\qty{30}{Hz}italic_δ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 2 italic_π × 30 italic_H italic_z to δf=2π×\qty570Hzsubscript𝛿𝑓2𝜋\qty570𝐻𝑧\delta_{f}=-2\pi\times\qty{570}{Hz}italic_δ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT = - 2 italic_π × 570 italic_H italic_z. (b) zθ𝑧𝜃z-\thetaitalic_z - italic_θ phase space diagram for ΩL=0.2ERPlanck-constant-over-2-pisubscriptΩL0.2subscript𝐸R\hbar\Omega_{\text{L}}=0.2\,E_{\text{R}}roman_ℏ roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT = 0.2 italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT and gn=0.25ER𝑔𝑛0.25subscript𝐸Rgn=0.25\,E_{\text{R}}italic_g italic_n = 0.25 italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT along with the experimental data. The points shown are obtained by calculating θ𝜃\thetaitalic_θ theoretically for the experimentally measured z𝑧zitalic_z values according to σz=a+bzdelimited-⟨⟩subscript𝜎𝑧𝑎𝑏𝑧\langle\sigma_{z}\rangle=a+bz⟨ italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ⟩ = italic_a + italic_b italic_z [[SeeSupplementaryMaterialsat][formoretheoreticaldetailsaboutthemodel, Josephson&GPdynamics, therelationwithBogoliubovspectrum, aswellassomeexperimentaldetailsabouttheparametersandBraggspectrum.]sup_mat]. Contour lines correspond to orbits with different energy.

i) Numerical simulations of GP Eq. (1), which agree well with the experimental data for all ΩLPlanck-constant-over-2-pisubscriptΩL\hbar\Omega_{\text{L}}roman_ℏ roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT.

ii) Numerical simulations of the effective dynamics (3) from the two-mode approximation. In such detuning quench-driven dynamics, we have θi=0subscript𝜃𝑖0\theta_{i}=0italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 0, as seen from the fixed point position in the phase space diagram. Starting from the initial zisubscript𝑧𝑖z_{i}italic_z start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT for δisubscript𝛿𝑖\delta_{i}italic_δ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, we numerically integrate Eq. (3) and determine the oscillation frequency.

iii) Perturbation analysis of the two-mode dynamics. For the case of a small quench, we can treat the dynamics around the fixed point (z0,θ0=0)subscript𝑧0subscript𝜃00(z_{0},\theta_{0}=0)( italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0 ) for δfsubscript𝛿𝑓\delta_{f}italic_δ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT as a perturbation, i.e., z=z0+δz𝑧subscript𝑧0𝛿𝑧z=z_{0}+\delta zitalic_z = italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_δ italic_z, θ=δθ𝜃𝛿𝜃\theta=\delta\thetaitalic_θ = italic_δ italic_θ, yielding

τδzsubscript𝜏𝛿𝑧\displaystyle\partial_{\tau}{\delta z}∂ start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT italic_δ italic_z =\displaystyle== 1z02δθ1superscriptsubscript𝑧02𝛿𝜃\displaystyle-\sqrt{1-z_{0}^{2}}~{}\delta\theta- square-root start_ARG 1 - italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG italic_δ italic_θ
τδθsubscript𝜏𝛿𝜃\displaystyle\partial_{\tau}{\delta\theta}∂ start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT italic_δ italic_θ =\displaystyle== [Λ+(1z02)3/2]δz,delimited-[]Λsuperscript1superscriptsubscript𝑧0232𝛿𝑧\displaystyle\left[\Lambda+(1-z_{0}^{2})^{-3/2}\right]~{}\delta z,[ roman_Λ + ( 1 - italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT - 3 / 2 end_POSTSUPERSCRIPT ] italic_δ italic_z , (4)

with initial conditions δz(0)=ziz0𝛿𝑧0subscript𝑧𝑖subscript𝑧0\delta z(0)=z_{i}-z_{0}italic_δ italic_z ( 0 ) = italic_z start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and δθ(0)=θiθ0=0𝛿𝜃0subscript𝜃𝑖subscript𝜃00\delta\theta(0)=\theta_{i}-\theta_{0}=0italic_δ italic_θ ( 0 ) = italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0. Therefore the oscillation frequency Δ=2K[1z02Λ+(1z02)1]1/2Δ2𝐾superscriptdelimited-[]1superscriptsubscript𝑧02Λsuperscript1superscriptsubscript𝑧02112\Delta=2K\left[\sqrt{1-z_{0}^{2}}~{}\Lambda+\left(1-z_{0}^{2}\right)^{-1}% \right]^{1/2}roman_Δ = 2 italic_K [ square-root start_ARG 1 - italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG roman_Λ + ( 1 - italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ] start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT. Note that the initial imbalance z0subscript𝑧0z_{0}italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT also depends on the lattice strength, and as ΩL0Planck-constant-over-2-pisubscriptΩL0\hbar\Omega_{\text{L}}\rightarrow 0roman_ℏ roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT → 0, one has z01subscript𝑧01z_{0}\rightarrow-1italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT → - 1 and K0𝐾0K\rightarrow 0italic_K → 0, leading to a finite ΔΔ\Deltaroman_Δ. A plot of z0subscript𝑧0z_{0}italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT with respect to ΩLPlanck-constant-over-2-pisubscriptΩL\hbar\Omega_{\text{L}}roman_ℏ roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT is shown in [[SeeSupplementaryMaterialsat][formoretheoreticaldetailsaboutthemodel, Josephson&GPdynamics, therelationwithBogoliubovspectrum, aswellassomeexperimentaldetailsabouttheparametersandBraggspectrum.]sup_mat]. The analytic expression agrees well with the numerical results in ii) but deviates from the experimental results and GP simulation significantly in the large ΩLPlanck-constant-over-2-pisubscriptΩL\hbar\Omega_{\text{L}}roman_ℏ roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT regime, where higher energy modes in the SOC band are coupled by the optical lattice, leading to the failure of the two-mode approximation.

iv) The zero quasimomentum gap of the Bogoliubov excitation spectrum. The sudden quench of the detuning leads to collective Bogoliubov quasiparticle excitations of the BEC located at quasimomentum q=0𝑞0q=0italic_q = 0 due to the lack of momentum transfer. When the quench is weak, only the lowest quasiparticle band is excited. Therefore, the plasma oscillation frequency can be determined from the lowest Bogoliubov band gap. Details of this analysis and its connection with the pseudo-Goldstone mode are discussed in the next section.

While the phase varies only within a small range for a Josephson plasma oscillation, it can continuously vary through [0,2π]02𝜋[0,2\pi][ 0 , 2 italic_π ] for a regular Josephson oscillation. We access this regime by choosing a larger quench δi=2π×\qty100Hzsubscript𝛿𝑖2𝜋\qty100𝐻𝑧\delta_{i}=2\pi\times\qty{100}{Hz}italic_δ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 2 italic_π × 100 italic_H italic_z and δf=2π×\qty500Hzsubscript𝛿𝑓2𝜋\qty500𝐻𝑧\delta_{f}=-2\pi\times\qty{500}{Hz}italic_δ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT = - 2 italic_π × 500 italic_H italic_z with ΩL=0.2ERPlanck-constant-over-2-pisubscriptΩL0.2subscript𝐸R\hbar\Omega_{\text{L}}=0.2\,E_{\text{R}}roman_ℏ roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT = 0.2 italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT so that the initial zisubscript𝑧𝑖z_{i}italic_z start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is far from the fixed point z0subscript𝑧0z_{0}italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. Fig. 4 (a) shows the spin-polarization oscillation as a function of the post-quench time t𝑡titalic_t, demonstrating a large amplitude of oscillation with a change of sign in σzdelimited-⟨⟩subscript𝜎𝑧\langle\sigma_{z}\rangle⟨ italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ⟩ and an oscillation frequency of 0.275±0.011ERplus-or-minus0.2750.011subscript𝐸R0.275\pm 0.011\,E_{\text{R}}0.275 ± 0.011 italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT. In Fig. 4 (b), we show a contour plot for the variation of the phase θ𝜃\thetaitalic_θ with z𝑧zitalic_z. The data points in Fig. 4 (b) are obtained by calculating θ𝜃\thetaitalic_θ theoretically for the experimentally measured z𝑧zitalic_z values according to σz=a+bzdelimited-⟨⟩subscript𝜎𝑧𝑎𝑏𝑧\langle\sigma_{z}\rangle=a+bz⟨ italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ⟩ = italic_a + italic_b italic_z [[SeeSupplementaryMaterialsat][formoretheoreticaldetailsaboutthemodel, Josephson&GPdynamics, therelationwithBogoliubovspectrum, aswellassomeexperimentaldetailsabouttheparametersandBraggspectrum.]sup_mat]. The error in z𝑧zitalic_z represents the standard error calculated from the standard deviation in σzdelimited-⟨⟩subscript𝜎𝑧\langle\sigma_{z}\rangle⟨ italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ⟩. Fig. 4 (a) shows the regular Josephson oscillation in momentum space, with a suboptimal agreement between the theory and experiment. The dark blue curve shows the numerical GP simulation results for a Raman detuning (δ𝛿\deltaitalic_δ) shift of \qty70Hz\qty70𝐻𝑧\qty{70}{Hz}70 italic_H italic_z, which is within the range of ±\qty100Hzplus-or-minus\qty100𝐻𝑧\pm\qty{100}{Hz}± 100 italic_H italic_z uncertainty and demonstrates a better agreement.

Connection with Bogoliubov spectrum and pseudo-Goldstone mode.

As discussed in iv), a small sudden Raman detuning quench generates collective excitations of the BEC from the ground band to the first excited band. The modified band structure with the incorporation of the optical lattice and mean field interaction is shown in Fig. 5 (a). Without the optical lattice and in the absence of a Raman detuning δ𝛿\deltaitalic_δ, the system has translational symmetry, and the interaction leads to two gapless Goldstone modes for the stripe phase [43]. The weak optical lattice breaks the translational symmetry explicitly, causing one Goldstone mode to become gapped at zero quasimomentum q=0𝑞0q=0italic_q = 0. This mode is then referred to as a pseudo-Goldstone mode, which is highly relevant in the context of particle and condensed matter physics [40]. The dependence of this pseudo-Goldstone gap on the mean-field interaction is shown in Fig. 5 (b). The comparison between the experimentally measured Josephson plasma oscillation frequency and the Bogoliubov excitation gap is shown in Figs. 3 (d) and (e). We see that the Bogoliubov gap agrees well with the experimental measurement as well as the GP simulation. Clearly, mean-field interactions play an important role in the zero quasimomentum band gap [[SeeSupplementaryMaterialsat][formoretheoreticaldetailsaboutthemodel, Josephson&GPdynamics, therelationwithBogoliubovspectrum, aswellassomeexperimentaldetailsabouttheparametersandBraggspectrum.]sup_mat]. As expected, the Bogoliubov gaps agree with those obtained from two-mode Josephson dynamics in shallow lattice regimes. The gap deviates significantly in deep lattice regimes due to the coupling with higher momentum modes.

Refer to caption
Figure 5: (a) Bogoliubov spectrum of H0subscript𝐻0H_{0}italic_H start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT for ΩR=2.7ERPlanck-constant-over-2-pisubscriptΩR2.7subscript𝐸R\hbar\Omega_{\text{R}}=2.7\,E_{\text{R}}roman_ℏ roman_Ω start_POSTSUBSCRIPT R end_POSTSUBSCRIPT = 2.7 italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT, ΩL=1ERPlanck-constant-over-2-pisubscriptΩL1subscript𝐸R\hbar\Omega_{\text{L}}=1\,E_{\text{R}}roman_ℏ roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT = 1 italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT, δ=2π×\qty500Hz𝛿2𝜋\qty500𝐻𝑧\delta=2\pi\times\qty{500}{Hz}italic_δ = 2 italic_π × 500 italic_H italic_z, and gn=0.25ER𝑔𝑛0.25subscript𝐸Rgn=0.25\,E_{\text{R}}italic_g italic_n = 0.25 italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT. (b) Dependence of zero quasimomentum (q=0𝑞0q=0italic_q = 0) band gap (ΔΔ\Deltaroman_Δ) on gn𝑔𝑛gnitalic_g italic_n for ΩL=0Planck-constant-over-2-pisubscriptΩL0\hbar\Omega_{\text{L}}=0roman_ℏ roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT = 0 (solid) and ΩL=1ERPlanck-constant-over-2-pisubscriptΩL1subscript𝐸R\hbar\Omega_{\text{L}}=1\,E_{\text{R}}roman_ℏ roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT = 1 italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT (dashed).

The connection between the Josephson oscillation frequency and the pseudo-Goldstone gap can be understood as follows. Consider a small deviation of the wave function away from the ground state

ψ=[(ϕl0+δϕl)χleikLx+(ϕr0+δϕr)χreikLx]eikbx,𝜓delimited-[]superscriptsubscriptitalic-ϕ𝑙0𝛿subscriptitalic-ϕ𝑙subscript𝜒𝑙superscript𝑒𝑖subscript𝑘L𝑥superscriptsubscriptitalic-ϕ𝑟0𝛿subscriptitalic-ϕ𝑟subscript𝜒𝑟superscript𝑒𝑖subscript𝑘L𝑥superscript𝑒𝑖subscript𝑘b𝑥\psi=\left[(\phi_{l}^{0}+\delta\phi_{l})\chi_{l}e^{-ik_{\text{L}}x}+(\phi_{r}^% {0}+\delta\phi_{r})\chi_{r}e^{ik_{\text{L}}x}\right]e^{ik_{\text{b}}x},italic_ψ = [ ( italic_ϕ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT + italic_δ italic_ϕ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ) italic_χ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_i italic_k start_POSTSUBSCRIPT L end_POSTSUBSCRIPT italic_x end_POSTSUPERSCRIPT + ( italic_ϕ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT + italic_δ italic_ϕ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ) italic_χ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_k start_POSTSUBSCRIPT L end_POSTSUBSCRIPT italic_x end_POSTSUPERSCRIPT ] italic_e start_POSTSUPERSCRIPT italic_i italic_k start_POSTSUBSCRIPT b end_POSTSUBSCRIPT italic_x end_POSTSUPERSCRIPT , (5)

that is, Bogoliubov excitations with the two-mode approximation. δϕl𝛿subscriptitalic-ϕ𝑙\delta\phi_{l}italic_δ italic_ϕ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT and δϕr𝛿subscriptitalic-ϕ𝑟\delta\phi_{r}italic_δ italic_ϕ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT can be obtained by solving their corresponding Bogoliubov equation [[SeeSupplementaryMaterialsat][formoretheoreticaldetailsaboutthemodel, Josephson&GPdynamics, therelationwithBogoliubovspectrum, aswellassomeexperimentaldetailsabouttheparametersandBraggspectrum.]sup_mat, 44]. Comparing with Eq. (4), we find δz=2jRe[(1)jδϕjϕj0]𝛿𝑧2subscript𝑗Redelimited-[]superscript1𝑗𝛿superscriptsubscriptitalic-ϕ𝑗superscriptsubscriptitalic-ϕ𝑗0\delta z=-2\sum_{j}\text{Re}[(-1)^{j}\delta\phi_{j}^{\ast}\phi_{j}^{0}]italic_δ italic_z = - 2 ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT Re [ ( - 1 ) start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT italic_δ italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT ] and δθ=jIm[(1)jδϕj/ϕj0]𝛿𝜃subscript𝑗Imdelimited-[]superscript1𝑗𝛿subscriptitalic-ϕ𝑗superscriptsubscriptitalic-ϕ𝑗0\delta\theta=-\sum_{j}\text{Im}[(-1)^{j}\delta\phi_{j}/\phi_{j}^{0}]italic_δ italic_θ = - ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT Im [ ( - 1 ) start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT italic_δ italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT / italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT ]. By examining the two lowest modes at q=0𝑞0q=0italic_q = 0, one finds that the gapped (gapless) one gives rise to non-vanishing (vanishing) δz𝛿𝑧\delta zitalic_δ italic_z and δθ𝛿𝜃\delta\thetaitalic_δ italic_θ. Therefore, the Josephson plasma oscillation frequency is just the zero quasimomentum Bogoliubov roton gap.

The symmetry-breaking origin of the pseudo-Goldstone mode can also be intuitively understood in the effective two-mode dynamics. Without the coupling K𝐾Kitalic_K (induced by the optical lattice), Eq. 3 has a Us(1)×Ua(1)subscript𝑈𝑠1subscript𝑈𝑎1U_{s}(1)\times U_{a}(1)italic_U start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( 1 ) × italic_U start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ( 1 ) symmetry, where Us(1)subscript𝑈𝑠1U_{s}(1)italic_U start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( 1 ) corresponds to the simultaneous rotation of two modes and Ua(1)subscript𝑈𝑎1U_{a}(1)italic_U start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ( 1 ) represents equal but opposite phase rotation of two modes. The spontaneous symmetry breaking leads to two uncoupled gapless modes (i.e., the Goldstone modes). The introduction of a coupling K0𝐾0K\neq 0italic_K ≠ 0 breaks the symmetry Ua(1)subscript𝑈𝑎1U_{a}(1)italic_U start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ( 1 ) and reduces the system symmetry to Us(1)subscript𝑈𝑠1U_{s}(1)italic_U start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( 1 ). This Us(1)subscript𝑈𝑠1U_{s}(1)italic_U start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( 1 ) symmetry is spontaneously broken, and the attendant Goldstone boson is absorbed and removed from the spectra via the Higgs mechanism [45]. Only the second Goldstone boson corresponding to symmetry Ua(1)subscript𝑈𝑎1U_{a}(1)italic_U start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ( 1 ) appears. The parameter K𝐾Kitalic_K is the soft breaking parameter of the symmetry Ua(1)subscript𝑈𝑎1U_{a}(1)italic_U start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ( 1 ), and the corresponding excitations become pseudo-Goldstone Bosons.

Conclusion & discussion.

Our work offers a new experimental platform for designing exotic quantum matter and engineering quantum simulators utilizing momentum states as a synthetic degree of freedom. For instance, applying Bragg scattering to the superfluid stripe ground state, one can measure the Bogoliubov excitation spectrum at finite quasimomentum, leading to the full characterization of the long-sought pseudo-Goldstone mode. Because of the spin-momentum coupling, the density interaction in the superfluid stripe phase could induce strong spin squeezing, which may be realized in our platform and used for quantum sensing. Applying small but periodic modulations of the Raman detuning can lead to the observation of a Shapiro resonance in the momentum space Josephson junction. These novel quantum phenomena enabled by the momentum space Josephson junction could potentially be useful for quantum technologies.

Acknowledgements.
Acknowledgements: A.M., C.S., M.K.H.O., S.M., and P.E. acknowledge funding from NSF through Grant No. PHY-1912540. P.E. also acknowledges support through the Ralph G. Yount Distinguished Professorship at WSU. We acknowledge experimental support from Ethan Crowell during the initial stage of this project. C.Z. is supported by AFOSR (FA9550-20-1-0220) and NSF (PHY-2409943, OMR-2228725, ECC-2411394). X.L. acknowledges support from the Innovation Program for Quantum Science and Technology (Grant No. 2021ZD0301200), the National Natural Science Foundation of China (No. 12275203), and the USTC start-up funding.

I Supplementary Materials

Two-mode model and Josephson dynamics.

To introduce the two-mode model, we begin by writing the wavefunction of the spinor BEC in the following form

ψ(x,t)=jϕj(t)χjei(2j1)kLx+ikbx,𝜓𝑥𝑡subscript𝑗subscriptitalic-ϕ𝑗𝑡subscript𝜒𝑗superscript𝑒𝑖2𝑗1subscript𝑘L𝑥𝑖subscript𝑘b𝑥\psi(x,t)=\sum_{j}\phi_{j}(t)\chi_{j}e^{i(2j-1)k_{\text{L}}x+ik_{\text{b}}x},italic_ψ ( italic_x , italic_t ) = ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t ) italic_χ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i ( 2 italic_j - 1 ) italic_k start_POSTSUBSCRIPT L end_POSTSUBSCRIPT italic_x + italic_i italic_k start_POSTSUBSCRIPT b end_POSTSUBSCRIPT italic_x end_POSTSUPERSCRIPT , (6)

where the integer j=J,,J+1𝑗𝐽𝐽1j=-J,...,J+1italic_j = - italic_J , … , italic_J + 1 represent the reciprocal lattice vectors and J𝐽Jitalic_J is the cutoff of the plane-wave modes. χjsubscript𝜒𝑗\chi_{j}italic_χ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT is the expansion spinor and ϕj(t)subscriptitalic-ϕ𝑗𝑡\phi_{j}(t)italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t ) are the coefficients. kbsubscript𝑘bk_{\text{b}}italic_k start_POSTSUBSCRIPT b end_POSTSUBSCRIPT is the bias momentum induced by detuning δ𝛿\deltaitalic_δ. The dynamics are governed by

itψ(x,t)=H0ψ(x,t)+g|ψ(x,t)|2ψ(x,t)𝑖subscript𝑡𝜓𝑥𝑡subscript𝐻0𝜓𝑥𝑡𝑔superscript𝜓𝑥𝑡2𝜓𝑥𝑡i\partial_{t}\psi(x,t)=H_{0}\psi(x,t)+g|\psi(x,t)|^{2}\psi(x,t)italic_i ∂ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_ψ ( italic_x , italic_t ) = italic_H start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ψ ( italic_x , italic_t ) + italic_g | italic_ψ ( italic_x , italic_t ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ψ ( italic_x , italic_t ) (7)

with g𝑔gitalic_g as the interaction strength.

We are interested in the low energy dynamics with sufficiently weak lattice and interaction strength, and thus make a two-mode approximation, that is, kee** only two modes at two band minima with j=0,1𝑗01j=0,1italic_j = 0 , 1 and treating the lattice and interactions as perturbations. Notice that j=0𝑗0j=0italic_j = 0 (j=1𝑗1j=1italic_j = 1) corresponds to the left (right) minimum labeled by l𝑙litalic_l (r𝑟ritalic_r) in the main text. Under the two-mode approximation, we will use j=l𝑗𝑙j=litalic_j = italic_l and j=r𝑗𝑟j=ritalic_j = italic_r to represent j=0𝑗0j=0italic_j = 0 and j=1𝑗1j=1italic_j = 1. The lattice couples the two modes while the interactions induce effective attractive interactions for each mode. First, we solve for the two band minima with ΩL=0subscriptΩL0\Omega_{\text{L}}=0roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT = 0 to obtain kbsubscript𝑘bk_{\text{b}}italic_k start_POSTSUBSCRIPT b end_POSTSUBSCRIPT, kLsubscript𝑘𝐿k_{L}italic_k start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT, χlsubscript𝜒𝑙\chi_{l}italic_χ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT, and χrsubscript𝜒𝑟\chi_{r}italic_χ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT. Then, we rewrite the dynamic equation by substituting Eq. 6 into Eq. 7

itϕl(t)𝑖subscript𝑡subscriptitalic-ϕ𝑙𝑡\displaystyle i\partial_{t}\phi_{l}(t)italic_i ∂ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_ϕ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ( italic_t ) =\displaystyle== (ElUnl)ϕl(t)Kϕr(t)subscript𝐸𝑙𝑈subscript𝑛𝑙subscriptitalic-ϕ𝑙𝑡𝐾subscriptitalic-ϕ𝑟𝑡\displaystyle(E_{l}-Un_{l})\phi_{l}(t)-K\phi_{r}(t)( italic_E start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT - italic_U italic_n start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ) italic_ϕ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ( italic_t ) - italic_K italic_ϕ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ( italic_t )
itϕr(t)𝑖subscript𝑡subscriptitalic-ϕ𝑟𝑡\displaystyle i\partial_{t}\phi_{r}(t)italic_i ∂ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_ϕ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ( italic_t ) =\displaystyle== (ErUnr)ϕr(t)Kϕl(t)subscript𝐸𝑟𝑈subscript𝑛𝑟subscriptitalic-ϕ𝑟𝑡superscript𝐾subscriptitalic-ϕ𝑙𝑡\displaystyle(E_{r}-Un_{r})\phi_{r}(t)-K^{\ast}\phi_{l}(t)( italic_E start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT - italic_U italic_n start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ) italic_ϕ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ( italic_t ) - italic_K start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT italic_ϕ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ( italic_t ) (8)

where Ej=𝑑xχjH0χj+(g+U)nsubscript𝐸𝑗differential-d𝑥superscriptsubscript𝜒𝑗subscript𝐻0subscript𝜒𝑗𝑔𝑈𝑛E_{j}=\int dx\chi_{j}^{\ast}H_{0}\chi_{j}+(g+U)nitalic_E start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = ∫ italic_d italic_x italic_χ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT italic_H start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_χ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT + ( italic_g + italic_U ) italic_n, U=g|χlχr|2𝑈𝑔superscriptsuperscriptsubscript𝜒𝑙subscript𝜒𝑟2U=g|\chi_{l}^{\ast}\chi_{r}|^{2}italic_U = italic_g | italic_χ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT italic_χ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT and K=ΩL2χlχr𝐾subscriptΩL2superscriptsubscript𝜒𝑙subscript𝜒𝑟K=\frac{\Omega_{\text{L}}}{2}\chi_{l}^{\ast}\chi_{r}italic_K = divide start_ARG roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG italic_χ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT italic_χ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT (we set K𝐾Kitalic_K to be real without loss of generality), with n=nl+nr𝑛subscript𝑛𝑙subscript𝑛𝑟n=n_{l}+n_{r}italic_n = italic_n start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT + italic_n start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT a constant and nj=|ϕj|2subscript𝑛𝑗superscriptsubscriptitalic-ϕ𝑗2n_{j}=|\phi_{j}|^{2}italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = | italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. The above equation is the Bose Josephson junction tunneling equation. We can write the wave function as ϕj=njeiθjsubscriptitalic-ϕ𝑗subscript𝑛𝑗superscript𝑒𝑖subscript𝜃𝑗\phi_{j}=\sqrt{n_{j}}e^{i\theta_{j}}italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = square-root start_ARG italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG italic_e start_POSTSUPERSCRIPT italic_i italic_θ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUPERSCRIPT, and in terms of the phase difference θ=θlθr𝜃subscript𝜃𝑙subscript𝜃𝑟\theta=\theta_{l}-\theta_{r}italic_θ = italic_θ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT - italic_θ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT and population difference z=nrnln𝑧subscript𝑛𝑟subscript𝑛𝑙𝑛z=\frac{n_{r}-n_{l}}{n}italic_z = divide start_ARG italic_n start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT - italic_n start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT end_ARG start_ARG italic_n end_ARG, Eq. (8) becomes

τzsubscript𝜏𝑧\displaystyle\partial_{\tau}{z}∂ start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT italic_z =\displaystyle== 1z2sinθ1superscript𝑧2𝜃\displaystyle-\sqrt{1-z^{2}}\sin{\theta}- square-root start_ARG 1 - italic_z start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG roman_sin italic_θ
τθsubscript𝜏𝜃\displaystyle\partial_{\tau}{\theta}∂ start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT italic_θ =\displaystyle== Λz+z1z2cosθΔEΛ𝑧𝑧1superscript𝑧2𝜃Δ𝐸\displaystyle\Lambda z+\frac{z}{\sqrt{1-z^{2}}}\cos{\theta}-\Delta Eroman_Λ italic_z + divide start_ARG italic_z end_ARG start_ARG square-root start_ARG 1 - italic_z start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_ARG roman_cos italic_θ - roman_Δ italic_E (9)

with τ=2Kt𝜏2𝐾𝑡\tau=2Ktitalic_τ = 2 italic_K italic_t, ΔE=(ElEr)/(2K)Δ𝐸subscript𝐸𝑙subscript𝐸𝑟2𝐾\Delta E=(E_{l}-E_{r})/(2K)roman_Δ italic_E = ( italic_E start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ) / ( 2 italic_K ) and Λ=Un/(2K)Λ𝑈𝑛2𝐾\Lambda=-Un/(2K)roman_Λ = - italic_U italic_n / ( 2 italic_K ). The effective Hamiltonian reads

Heff=Λ2z21z2cosθΔEz.subscript𝐻effΛ2superscript𝑧21superscript𝑧2𝜃Δ𝐸𝑧H_{\text{eff}}=\frac{\Lambda}{2}z^{2}-\sqrt{1-z^{2}}\cos{\theta}-\Delta Ez.italic_H start_POSTSUBSCRIPT eff end_POSTSUBSCRIPT = divide start_ARG roman_Λ end_ARG start_ARG 2 end_ARG italic_z start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - square-root start_ARG 1 - italic_z start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG roman_cos italic_θ - roman_Δ italic_E italic_z . (10)

The ground state (z0,θ0subscript𝑧0subscript𝜃0z_{0},\theta_{0}italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT) can be obtained by finding the minima of Heffsubscript𝐻effH_{\text{eff}}italic_H start_POSTSUBSCRIPT eff end_POSTSUBSCRIPT. The variation of z0subscript𝑧0z_{0}italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT with respect to ΩLPlanck-constant-over-2-pisubscriptΩL\hbar\Omega_{\text{L}}roman_ℏ roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT is shown in Fig. 6. Without the coupling K𝐾Kitalic_K, the two-mode system Eq. 8 has a Us(1)×Ua(1)subscript𝑈𝑠1subscript𝑈𝑎1U_{s}(1)\times U_{a}(1)italic_U start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( 1 ) × italic_U start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ( 1 ) symmetry. The spontaneous symmetry breaking leads to two uncoupled gapless modes (i.e., the Nambu-Goldstone modes). The introduction of a coupling K0𝐾0K\neq 0italic_K ≠ 0 breaks the symmetry Ua(1)subscript𝑈𝑎1U_{a}(1)italic_U start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ( 1 ) and reduces the system symmetry to Us(1)subscript𝑈𝑠1U_{s}(1)italic_U start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( 1 ) (i.e., simultaneous rotation of the two modes). This Us(1)subscript𝑈𝑠1U_{s}(1)italic_U start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( 1 ) symmetry is spontaneously broken, and the attendant Nambu-Goldstone boson is absorbed and removed from the spectra via the Higgs mechanism [45]. Only the second Nambu-Goldstone boson corresponding to symmetry Ua(1)subscript𝑈𝑎1U_{a}(1)italic_U start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ( 1 ) (i.e., equal but opposite phase rotation of the two modes) appears. The parameter K𝐾Kitalic_K is the soft breaking parameter of the symmetry Ua(1)subscript𝑈𝑎1U_{a}(1)italic_U start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ( 1 ), and the corresponding excitations become pseudo Nambu-Goldstone Bosons. The frequency associated with these excitations is correspondingly small.

Refer to caption
Figure 6: Initial imbalance (z0subscript𝑧0z_{0}italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT) as a function of the optical lattice coupling strength (ΩLPlanck-constant-over-2-pisubscriptΩL\hbar\Omega_{\text{L}}roman_ℏ roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT). All other parameters are the same as those in Fig. 3 in the main text.

Treatment of quench dynamics using the GP equation.

Consider the spinor wave function ψ(x,t)=[ψ(x,t),ψ(x,t)]T𝜓𝑥𝑡superscriptsubscript𝜓𝑥𝑡subscript𝜓𝑥𝑡𝑇\psi(x,t)=[\psi_{\uparrow}(x,t),\psi_{\downarrow}(x,t)]^{T}italic_ψ ( italic_x , italic_t ) = [ italic_ψ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT ( italic_x , italic_t ) , italic_ψ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT ( italic_x , italic_t ) ] start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT. The dynamics are characterized by the Gross-Pitaevskii (GP) equation

iψ(x,t)t=(H0+Vint)ψ(x,t),𝑖𝜓𝑥𝑡𝑡subscript𝐻0subscript𝑉int𝜓𝑥𝑡i\frac{\partial\psi(x,t)}{\partial t}=(H_{0}+V_{\text{int}})\psi(x,t),italic_i divide start_ARG ∂ italic_ψ ( italic_x , italic_t ) end_ARG start_ARG ∂ italic_t end_ARG = ( italic_H start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_V start_POSTSUBSCRIPT int end_POSTSUBSCRIPT ) italic_ψ ( italic_x , italic_t ) , (11)

with

Vint=(g|ψ|2+g|ψ|200g|ψ|2+g|ψ|2).subscript𝑉intsubscript𝑔absentsuperscriptsubscript𝜓2subscript𝑔absentsuperscriptsubscript𝜓200subscript𝑔absentsuperscriptsubscript𝜓2subscript𝑔absentsuperscriptsubscript𝜓2V_{\text{int}}=\left(\begin{array}[]{cc}g_{\uparrow\uparrow}|\psi_{\uparrow}|^% {2}+g_{\uparrow\downarrow}|\psi_{\downarrow}|^{2}&0\\ 0&g_{\downarrow\downarrow}|\psi_{\downarrow}|^{2}+g_{\uparrow\downarrow}|\psi_% {\uparrow}|^{2}\end{array}\right).italic_V start_POSTSUBSCRIPT int end_POSTSUBSCRIPT = ( start_ARRAY start_ROW start_CELL italic_g start_POSTSUBSCRIPT ↑ ↑ end_POSTSUBSCRIPT | italic_ψ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_g start_POSTSUBSCRIPT ↑ ↓ end_POSTSUBSCRIPT | italic_ψ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_CELL start_CELL 0 end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL italic_g start_POSTSUBSCRIPT ↓ ↓ end_POSTSUBSCRIPT | italic_ψ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_g start_POSTSUBSCRIPT ↑ ↓ end_POSTSUBSCRIPT | italic_ψ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_CELL end_ROW end_ARRAY ) . (12)

Here, gsssubscript𝑔𝑠superscript𝑠g_{ss^{\prime}}italic_g start_POSTSUBSCRIPT italic_s italic_s start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT denotes the interaction strength between atoms in s𝑠sitalic_s and ssuperscript𝑠s^{\prime}italic_s start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT states. We can adopt the isotropic approximation for the interaction and set g=g=g=gsubscript𝑔absentsubscript𝑔absentsubscript𝑔absent𝑔g_{\uparrow\uparrow}=g_{\downarrow\downarrow}=g_{\uparrow\downarrow}=gitalic_g start_POSTSUBSCRIPT ↑ ↑ end_POSTSUBSCRIPT = italic_g start_POSTSUBSCRIPT ↓ ↓ end_POSTSUBSCRIPT = italic_g start_POSTSUBSCRIPT ↑ ↓ end_POSTSUBSCRIPT = italic_g, which is an excellent approximation for 87Rb atoms. Before the quench, the system is in the ground state, which can be obtained by the imaginary time evolution of the GP equation. With this ground state as the initial state, the quench dynamics (i.e., the polarization oscillation) can be obtained by solving the real-time GP equation.

Bogoliubov spectrum.

To connect the Josephson oscillation frequency with the Bogoliubov spectrum, we first calculate the spectrum of Bogoliubov excitations. We write the deviations of the wavefunctions with respect to the ground state as

ψs(x,t)=eiμt[ψ0s(x)+δψs(x,t)]subscript𝜓𝑠𝑥𝑡superscript𝑒𝑖𝜇𝑡delimited-[]subscript𝜓0𝑠𝑥𝛿subscript𝜓𝑠𝑥𝑡\psi_{s}(x,t)=e^{-i\mu t}\left[\psi_{0s}(x)+\delta\psi_{s}(x,t)\right]italic_ψ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_x , italic_t ) = italic_e start_POSTSUPERSCRIPT - italic_i italic_μ italic_t end_POSTSUPERSCRIPT [ italic_ψ start_POSTSUBSCRIPT 0 italic_s end_POSTSUBSCRIPT ( italic_x ) + italic_δ italic_ψ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_x , italic_t ) ] (13)

where ψ0s(x)subscript𝜓0𝑠𝑥\psi_{0s}(x)italic_ψ start_POSTSUBSCRIPT 0 italic_s end_POSTSUBSCRIPT ( italic_x ) is the ground state, and the deviations take the form δψs(x,t)=us(x)eiεt+vs(x)eiεt𝛿subscript𝜓𝑠𝑥𝑡subscript𝑢𝑠𝑥superscript𝑒𝑖𝜀𝑡superscriptsubscript𝑣𝑠𝑥superscript𝑒𝑖𝜀𝑡\delta\psi_{s}(x,t)=u_{s}(x)e^{-i\varepsilon t}+v_{s}^{\ast}(x)e^{i\varepsilon t}italic_δ italic_ψ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_x , italic_t ) = italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_x ) italic_e start_POSTSUPERSCRIPT - italic_i italic_ε italic_t end_POSTSUPERSCRIPT + italic_v start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( italic_x ) italic_e start_POSTSUPERSCRIPT italic_i italic_ε italic_t end_POSTSUPERSCRIPT. The amplitudes us(x)subscript𝑢𝑠𝑥u_{s}(x)italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_x ) and vs(x)subscript𝑣𝑠𝑥v_{s}(x)italic_v start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_x ) satisfy the normalization condition s0d𝑑x[|us(x)|2|vs(x)|2]=1subscript𝑠superscriptsubscript0𝑑differential-d𝑥delimited-[]superscriptsubscript𝑢𝑠𝑥2superscriptsubscript𝑣𝑠𝑥21\sum_{s}\int_{0}^{d}dx[|u_{s}(x)|^{2}-|v_{s}(x)|^{2}]=1∑ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT italic_d italic_x [ | italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_x ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - | italic_v start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_x ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] = 1, with d=2π/kL𝑑2𝜋subscript𝑘Ld=2\pi/k_{\text{L}}italic_d = 2 italic_π / italic_k start_POSTSUBSCRIPT L end_POSTSUBSCRIPT being the lattice period and μ𝜇\muitalic_μ the chemical potential. Substituting Eq. (13) into Eq. (11), we obtain the Bogoliubov equation as [u,u,v,v]T=ε[u,u,v,v]Tsuperscriptsubscript𝑢subscript𝑢subscript𝑣subscript𝑣𝑇𝜀superscriptsubscript𝑢subscript𝑢subscript𝑣subscript𝑣𝑇\mathcal{H}[u_{\uparrow},u_{\downarrow},v_{\uparrow},v_{\downarrow}]^{T}=% \varepsilon[u_{\uparrow},u_{\downarrow},v_{\uparrow},v_{\downarrow}]^{T}caligraphic_H [ italic_u start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT , italic_u start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT , italic_v start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT , italic_v start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT ] start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT = italic_ε [ italic_u start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT , italic_u start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT , italic_v start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT , italic_v start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT ] start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT, where the Bogoliubov Hamiltonian is given by

=(ΩR2+gψ0ψ0gψ02gψ0ψ0ΩR2+gψ0ψ0gψ0ψ0gψ02gψ02gψ0ψ0ΩR2gψ0ψ0gψ0ψ0gψ02ΩR2gψ0ψ0),subscriptsubscriptΩR2subscript𝑔absentsubscript𝜓0absentsuperscriptsubscript𝜓0absent𝑔superscriptsubscript𝜓0absent2subscript𝑔absentsubscript𝜓0absentsubscript𝜓0absentsubscriptΩR2subscript𝑔absentsuperscriptsubscript𝜓0absentsubscript𝜓0absentsubscriptsubscript𝑔absentsubscript𝜓0absentsubscript𝜓0absent𝑔superscriptsubscript𝜓0absent2𝑔superscriptsubscript𝜓0absentabsent2subscript𝑔absentsuperscriptsubscript𝜓0absentsuperscriptsubscript𝜓0absentsuperscriptsubscriptsubscriptΩR2subscript𝑔absentsuperscriptsubscript𝜓0absentsubscript𝜓0absentsubscript𝑔absentsuperscriptsubscript𝜓0absentsuperscriptsubscript𝜓0absent𝑔superscriptsubscript𝜓0absentabsent2subscriptΩR2subscript𝑔absentsubscript𝜓0absentsuperscriptsubscript𝜓0absentsuperscriptsubscript\mathcal{H}=\left(\begin{array}[]{cccc}\mathcal{H}_{\uparrow}&\frac{\Omega_{% \text{R}}}{2}+g_{\uparrow\downarrow}\psi_{0\uparrow}\psi_{0\downarrow}^{\ast}&% g\psi_{0\uparrow}^{2}&g_{\uparrow\downarrow}\psi_{0\uparrow}\psi_{0\downarrow}% \\ \frac{\Omega_{\text{R}}}{2}+g_{\uparrow\downarrow}\psi_{0\uparrow}^{\ast}\psi_% {0\downarrow}&\mathcal{H}_{\downarrow}&g_{\uparrow\downarrow}\psi_{0\uparrow}% \psi_{0\downarrow}&g\psi_{0\downarrow}^{2}\\ -g\psi_{0\uparrow}^{\ast 2}&-g_{\uparrow\downarrow}\psi_{0\uparrow}^{\ast}\psi% _{0\downarrow}^{\ast}&-\mathcal{H}_{\uparrow}^{\ast}&-\frac{\Omega_{\text{R}}}% {2}-g_{\uparrow\downarrow}\psi_{0\uparrow}^{\ast}\psi_{0\downarrow}\\ -g_{\uparrow\downarrow}\psi_{0\uparrow}^{\ast}\psi_{0\downarrow}^{\ast}&-g\psi% _{0\downarrow}^{\ast 2}&-\frac{\Omega_{\text{R}}}{2}-g_{\uparrow\downarrow}% \psi_{0\uparrow}\psi_{0\downarrow}^{\ast}&-\mathcal{H}_{\downarrow}^{\ast}\end% {array}\right),caligraphic_H = ( start_ARRAY start_ROW start_CELL caligraphic_H start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT end_CELL start_CELL divide start_ARG roman_Ω start_POSTSUBSCRIPT R end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG + italic_g start_POSTSUBSCRIPT ↑ ↓ end_POSTSUBSCRIPT italic_ψ start_POSTSUBSCRIPT 0 ↑ end_POSTSUBSCRIPT italic_ψ start_POSTSUBSCRIPT 0 ↓ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT end_CELL start_CELL italic_g italic_ψ start_POSTSUBSCRIPT 0 ↑ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_CELL start_CELL italic_g start_POSTSUBSCRIPT ↑ ↓ end_POSTSUBSCRIPT italic_ψ start_POSTSUBSCRIPT 0 ↑ end_POSTSUBSCRIPT italic_ψ start_POSTSUBSCRIPT 0 ↓ end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL divide start_ARG roman_Ω start_POSTSUBSCRIPT R end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG + italic_g start_POSTSUBSCRIPT ↑ ↓ end_POSTSUBSCRIPT italic_ψ start_POSTSUBSCRIPT 0 ↑ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT italic_ψ start_POSTSUBSCRIPT 0 ↓ end_POSTSUBSCRIPT end_CELL start_CELL caligraphic_H start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT end_CELL start_CELL italic_g start_POSTSUBSCRIPT ↑ ↓ end_POSTSUBSCRIPT italic_ψ start_POSTSUBSCRIPT 0 ↑ end_POSTSUBSCRIPT italic_ψ start_POSTSUBSCRIPT 0 ↓ end_POSTSUBSCRIPT end_CELL start_CELL italic_g italic_ψ start_POSTSUBSCRIPT 0 ↓ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_CELL end_ROW start_ROW start_CELL - italic_g italic_ψ start_POSTSUBSCRIPT 0 ↑ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ 2 end_POSTSUPERSCRIPT end_CELL start_CELL - italic_g start_POSTSUBSCRIPT ↑ ↓ end_POSTSUBSCRIPT italic_ψ start_POSTSUBSCRIPT 0 ↑ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT italic_ψ start_POSTSUBSCRIPT 0 ↓ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT end_CELL start_CELL - caligraphic_H start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT end_CELL start_CELL - divide start_ARG roman_Ω start_POSTSUBSCRIPT R end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG - italic_g start_POSTSUBSCRIPT ↑ ↓ end_POSTSUBSCRIPT italic_ψ start_POSTSUBSCRIPT 0 ↑ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT italic_ψ start_POSTSUBSCRIPT 0 ↓ end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL - italic_g start_POSTSUBSCRIPT ↑ ↓ end_POSTSUBSCRIPT italic_ψ start_POSTSUBSCRIPT 0 ↑ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT italic_ψ start_POSTSUBSCRIPT 0 ↓ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT end_CELL start_CELL - italic_g italic_ψ start_POSTSUBSCRIPT 0 ↓ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ 2 end_POSTSUPERSCRIPT end_CELL start_CELL - divide start_ARG roman_Ω start_POSTSUBSCRIPT R end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG - italic_g start_POSTSUBSCRIPT ↑ ↓ end_POSTSUBSCRIPT italic_ψ start_POSTSUBSCRIPT 0 ↑ end_POSTSUBSCRIPT italic_ψ start_POSTSUBSCRIPT 0 ↓ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT end_CELL start_CELL - caligraphic_H start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT end_CELL end_ROW end_ARRAY ) , (14)

with

=2/x2+2i/xδ/2+VL(x)μ+2g|ψ0|2+g|ψ0|2,subscriptsuperscript2superscript𝑥22𝑖𝑥𝛿2subscript𝑉L𝑥𝜇2𝑔superscriptsubscript𝜓0absent2subscript𝑔absentsuperscriptsubscript𝜓0absent2\mathcal{H}_{\uparrow}=-\partial^{2}/\partial x^{2}+2i\partial/\partial x-% \delta/2+V_{\text{L}}(x)-\mu+2g|\psi_{0\uparrow}|^{2}+g_{\uparrow\downarrow}|% \psi_{0\downarrow}|^{2},caligraphic_H start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT = - ∂ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / ∂ italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 2 italic_i ∂ / ∂ italic_x - italic_δ / 2 + italic_V start_POSTSUBSCRIPT L end_POSTSUBSCRIPT ( italic_x ) - italic_μ + 2 italic_g | italic_ψ start_POSTSUBSCRIPT 0 ↑ end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_g start_POSTSUBSCRIPT ↑ ↓ end_POSTSUBSCRIPT | italic_ψ start_POSTSUBSCRIPT 0 ↓ end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , (15)
=2/x22i/x+δ/2+VL(x)μ+2g|ψ0|2+g|ψ0|2.subscriptsuperscript2superscript𝑥22𝑖𝑥𝛿2subscript𝑉L𝑥𝜇2𝑔superscriptsubscript𝜓0absent2subscript𝑔absentsuperscriptsubscript𝜓0absent2\mathcal{H}_{\downarrow}=-\partial^{2}/\partial x^{2}-2i\partial/\partial x+% \delta/2+V_{\text{L}}(x)-\mu+2g|\psi_{0\downarrow}|^{2}+g_{\uparrow\downarrow}% |\psi_{0\uparrow}|^{2}.caligraphic_H start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT = - ∂ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / ∂ italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 2 italic_i ∂ / ∂ italic_x + italic_δ / 2 + italic_V start_POSTSUBSCRIPT L end_POSTSUBSCRIPT ( italic_x ) - italic_μ + 2 italic_g | italic_ψ start_POSTSUBSCRIPT 0 ↓ end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_g start_POSTSUBSCRIPT ↑ ↓ end_POSTSUBSCRIPT | italic_ψ start_POSTSUBSCRIPT 0 ↑ end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT . (16)

The excitation spectra can be calculated numerically by expanding us(x)subscript𝑢𝑠𝑥u_{s}(x)italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_x ) and vs(x)subscript𝑣𝑠𝑥v_{s}(x)italic_v start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_x ) in the Bloch basis. Each excitation spectrum is periodic in momentum space with the Brillouin zone [0,2kL]02subscript𝑘L[0,2k_{\text{L}}][ 0 , 2 italic_k start_POSTSUBSCRIPT L end_POSTSUBSCRIPT ] determined by the lattice period.

The ground state takes the following form

ψ0(x)=jϕj0χjei(2j1)kLx+ikbx,subscript𝜓0𝑥subscript𝑗subscriptsuperscriptitalic-ϕ0𝑗subscript𝜒𝑗superscript𝑒𝑖2𝑗1subscript𝑘L𝑥𝑖subscript𝑘b𝑥\psi_{0}(x)=\sum_{j}\phi^{0}_{j}\chi_{j}e^{i(2j-1)k_{\text{L}}x+ik_{\text{b}}x},italic_ψ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_x ) = ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_ϕ start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_χ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i ( 2 italic_j - 1 ) italic_k start_POSTSUBSCRIPT L end_POSTSUBSCRIPT italic_x + italic_i italic_k start_POSTSUBSCRIPT b end_POSTSUBSCRIPT italic_x end_POSTSUPERSCRIPT , (17)

where the integer j=J,,J+1𝑗𝐽𝐽1j=-J,...,J+1italic_j = - italic_J , … , italic_J + 1 represent the reciprocal lattice vectors and J𝐽Jitalic_J is the cutoff of the plane-wave modes. χjsubscript𝜒𝑗\chi_{j}italic_χ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT is the expansion spinor and ϕj0subscriptsuperscriptitalic-ϕ0𝑗\phi^{0}_{j}italic_ϕ start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT are the coefficients. kbsubscript𝑘bk_{\text{b}}italic_k start_POSTSUBSCRIPT b end_POSTSUBSCRIPT is the bias momentum induced by the detuning δ𝛿\deltaitalic_δ. In the quasi-momentum frame, the two band minima become asymmetric with respect to the zero quasi-momentum if the detuning is non-zero. From the form of the ansatz, one can see that kbsubscript𝑘𝑏k_{b}italic_k start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT is located at the center of the two band minima. So when there is a non-zero detuning, kbsubscript𝑘𝑏k_{b}italic_k start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT is also nonzero [36].

The perturbation amplitudes us(x)subscript𝑢𝑠𝑥u_{s}(x)italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_x ) and vs(x)subscript𝑣𝑠𝑥v_{s}(x)italic_v start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_x ) are expanded in the Bloch form in terms of the reciprocal lattice vectors:

us(x)=m=MM+1Us,mei(kb+q)x+i(2m1)kLx,subscript𝑢𝑠𝑥superscriptsubscript𝑚𝑀𝑀1subscript𝑈𝑠𝑚superscript𝑒𝑖subscript𝑘b𝑞𝑥𝑖2𝑚1subscript𝑘L𝑥u_{s}(x)=\sum_{m=-M}^{M+1}U_{s,m}e^{i(k_{\text{b}}+q)x+i(2m-1)k_{\text{L}}x},italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_x ) = ∑ start_POSTSUBSCRIPT italic_m = - italic_M end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_M + 1 end_POSTSUPERSCRIPT italic_U start_POSTSUBSCRIPT italic_s , italic_m end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i ( italic_k start_POSTSUBSCRIPT b end_POSTSUBSCRIPT + italic_q ) italic_x + italic_i ( 2 italic_m - 1 ) italic_k start_POSTSUBSCRIPT L end_POSTSUBSCRIPT italic_x end_POSTSUPERSCRIPT , (18)
vs(x)=m=MM+1Vs,mei(kb+q)xi(2m1)kLx,subscript𝑣𝑠𝑥superscriptsubscript𝑚𝑀𝑀1subscript𝑉𝑠𝑚superscript𝑒𝑖subscript𝑘b𝑞𝑥𝑖2𝑚1subscript𝑘L𝑥v_{s}(x)=\sum_{m=-M}^{M+1}V_{s,m}e^{i(k_{\text{b}}+q)x-i(2m-1)k_{\text{L}}x},italic_v start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_x ) = ∑ start_POSTSUBSCRIPT italic_m = - italic_M end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_M + 1 end_POSTSUPERSCRIPT italic_V start_POSTSUBSCRIPT italic_s , italic_m end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i ( italic_k start_POSTSUBSCRIPT b end_POSTSUBSCRIPT + italic_q ) italic_x - italic_i ( 2 italic_m - 1 ) italic_k start_POSTSUBSCRIPT L end_POSTSUBSCRIPT italic_x end_POSTSUPERSCRIPT , (19)

where q𝑞qitalic_q is the Bloch wavevector of the excitations, and M𝑀Mitalic_M is the cutoff of the expansion. The whole spectrum can be obtained by substituting the expansions into the Bogoliubov equation.

The ground state of the system is a striped phase due to the presence of the optical lattice. The stripe is enforced by the optical lattice, and the translational symmetry (equivalent to the Ua(1)subscript𝑈𝑎1U_{a}(1)italic_U start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ( 1 ) symmetry) is explicitly broken by the lattice, which is different from the interaction-induced supersolid stripe phase where the translational symmetry is spontaneously broken. At zero-momentum, one of the two lowest Bogoliubov bands corresponds to a gapless Nambu-Goldstone mode (related to the spontaneous breaking of Us(1)subscript𝑈𝑠1U_{s}(1)italic_U start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( 1 ) symmetry), the other is a gapped pseudo-Nambu-Goldstone mode (related to the explicit breaking of Ua(1)subscript𝑈𝑎1U_{a}(1)italic_U start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ( 1 ) symmetry).

As mentioned above, without the lattice, the translational symmetry (i.e., Ua(1)subscript𝑈𝑎1U_{a}(1)italic_U start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ( 1 ) symmetry) may be spontaneously broken by interaction. For example, anti-ferromagnetic interaction (ggsubscript𝑔absentsubscript𝑔absentg_{\uparrow\downarrow}-g_{\uparrow\uparrow}italic_g start_POSTSUBSCRIPT ↑ ↓ end_POSTSUBSCRIPT - italic_g start_POSTSUBSCRIPT ↑ ↑ end_POSTSUBSCRIPT) may lead to the supersolid stripe phase [46, 47, 37, 48]; however, it is very weak, and the supersolid stripe phase only exits for extremely low values of ΩRsubscriptΩR\Omega_{\text{R}}roman_Ω start_POSTSUBSCRIPT R end_POSTSUBSCRIPT and δ𝛿\deltaitalic_δ, making it difficult to observe in an experiment. For the dynamics studied in this paper, the effect of such anti-ferromagnetic interaction is negligible, and we can safely adopt the isotropic approach by setting g=g=g=gsubscript𝑔absentsubscript𝑔absentsubscript𝑔absent𝑔g_{\uparrow\uparrow}=g_{\downarrow\downarrow}=g_{\uparrow\downarrow}=gitalic_g start_POSTSUBSCRIPT ↑ ↑ end_POSTSUBSCRIPT = italic_g start_POSTSUBSCRIPT ↓ ↓ end_POSTSUBSCRIPT = italic_g start_POSTSUBSCRIPT ↑ ↓ end_POSTSUBSCRIPT = italic_g.

We note that the calculation of the correct Bogoliubov spectrum requires including high momentum modes with large J,M𝐽𝑀J,Mitalic_J , italic_M. We have set J=9𝐽9J=9italic_J = 9 and M=7𝑀7M=7italic_M = 7 in calculating the Bogoliubov spectrum. To show that the Bogoliubov gap at q=0𝑞0q=0italic_q = 0 is connected with the Josephson junction oscillation frequency, we consider only the corresponding excitation. Initially (t=0𝑡0t=0italic_t = 0), the wave function of the excitation reads δψs(x,t=0)=us(x)+vs(x)𝛿subscript𝜓𝑠𝑥𝑡0subscript𝑢𝑠𝑥subscriptsuperscript𝑣𝑠𝑥\delta\psi_{s}(x,t=0)=u_{s}(x)+v^{*}_{s}(x)italic_δ italic_ψ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_x , italic_t = 0 ) = italic_u start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_x ) + italic_v start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_x ), and we examine the distributions at the two band minima. That is, we consider the ground state ψ0s(ψ0,lseikLx+ψ0,rseikLx)eikbxsimilar-to-or-equalssubscript𝜓0𝑠subscript𝜓0𝑙𝑠superscript𝑒𝑖subscript𝑘L𝑥subscript𝜓0𝑟𝑠superscript𝑒𝑖subscript𝑘L𝑥superscript𝑒𝑖subscript𝑘b𝑥\psi_{0s}\simeq(\psi_{0,ls}e^{-ik_{\text{L}}x}+\psi_{0,rs}e^{ik_{\text{L}}x})e% ^{ik_{\text{b}}x}italic_ψ start_POSTSUBSCRIPT 0 italic_s end_POSTSUBSCRIPT ≃ ( italic_ψ start_POSTSUBSCRIPT 0 , italic_l italic_s end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_i italic_k start_POSTSUBSCRIPT L end_POSTSUBSCRIPT italic_x end_POSTSUPERSCRIPT + italic_ψ start_POSTSUBSCRIPT 0 , italic_r italic_s end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_k start_POSTSUBSCRIPT L end_POSTSUBSCRIPT italic_x end_POSTSUPERSCRIPT ) italic_e start_POSTSUPERSCRIPT italic_i italic_k start_POSTSUBSCRIPT b end_POSTSUBSCRIPT italic_x end_POSTSUPERSCRIPT and the excitation δψs(δψlseikLx+δψrseikLx)eikbxsimilar-to-or-equals𝛿subscript𝜓𝑠𝛿subscript𝜓𝑙𝑠superscript𝑒𝑖subscript𝑘L𝑥𝛿subscript𝜓𝑟𝑠superscript𝑒𝑖subscript𝑘L𝑥superscript𝑒𝑖subscript𝑘b𝑥\delta\psi_{s}\simeq(\delta\psi_{ls}e^{-ik_{\text{L}}x}+\delta\psi_{rs}e^{ik_{% \text{L}}x})e^{ik_{\text{b}}x}italic_δ italic_ψ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ≃ ( italic_δ italic_ψ start_POSTSUBSCRIPT italic_l italic_s end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_i italic_k start_POSTSUBSCRIPT L end_POSTSUBSCRIPT italic_x end_POSTSUPERSCRIPT + italic_δ italic_ψ start_POSTSUBSCRIPT italic_r italic_s end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_k start_POSTSUBSCRIPT L end_POSTSUBSCRIPT italic_x end_POSTSUPERSCRIPT ) italic_e start_POSTSUPERSCRIPT italic_i italic_k start_POSTSUBSCRIPT b end_POSTSUBSCRIPT italic_x end_POSTSUPERSCRIPT. The distribution amplitudes ψ0,jssubscript𝜓0𝑗𝑠\psi_{0,js}italic_ψ start_POSTSUBSCRIPT 0 , italic_j italic_s end_POSTSUBSCRIPT and δψjs𝛿subscript𝜓𝑗𝑠\delta\psi_{js}italic_δ italic_ψ start_POSTSUBSCRIPT italic_j italic_s end_POSTSUBSCRIPT can be real under a proper gauge choice and are shown in Fig. 7. We find that δψlsψ0,lsproportional-to𝛿subscript𝜓𝑙𝑠subscript𝜓0𝑙𝑠\delta\psi_{ls}\propto-\psi_{0,ls}italic_δ italic_ψ start_POSTSUBSCRIPT italic_l italic_s end_POSTSUBSCRIPT ∝ - italic_ψ start_POSTSUBSCRIPT 0 , italic_l italic_s end_POSTSUBSCRIPT and δψrsψ0,rsproportional-to𝛿subscript𝜓𝑟𝑠subscript𝜓0𝑟𝑠\delta\psi_{rs}\propto\psi_{0,rs}italic_δ italic_ψ start_POSTSUBSCRIPT italic_r italic_s end_POSTSUBSCRIPT ∝ italic_ψ start_POSTSUBSCRIPT 0 , italic_r italic_s end_POSTSUBSCRIPT, so we can write the wave function as

ψ=[(ϕl0+δϕl)χleikLx+(ϕr0+δϕr)χreikLx]eikbx,𝜓delimited-[]superscriptsubscriptitalic-ϕ𝑙0𝛿subscriptitalic-ϕ𝑙subscript𝜒𝑙superscript𝑒𝑖subscript𝑘L𝑥superscriptsubscriptitalic-ϕ𝑟0𝛿subscriptitalic-ϕ𝑟subscript𝜒𝑟superscript𝑒𝑖subscript𝑘L𝑥superscript𝑒𝑖subscript𝑘b𝑥\psi=\left[(\phi_{l}^{0}+\delta\phi_{l})\chi_{l}e^{-ik_{\text{L}}x}+(\phi_{r}^% {0}+\delta\phi_{r})\chi_{r}e^{ik_{\text{L}}x}\right]e^{ik_{\text{b}}x},italic_ψ = [ ( italic_ϕ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT + italic_δ italic_ϕ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ) italic_χ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_i italic_k start_POSTSUBSCRIPT L end_POSTSUBSCRIPT italic_x end_POSTSUPERSCRIPT + ( italic_ϕ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT + italic_δ italic_ϕ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ) italic_χ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_k start_POSTSUBSCRIPT L end_POSTSUBSCRIPT italic_x end_POSTSUPERSCRIPT ] italic_e start_POSTSUPERSCRIPT italic_i italic_k start_POSTSUBSCRIPT b end_POSTSUBSCRIPT italic_x end_POSTSUPERSCRIPT ,

and initially (t=0𝑡0t=0italic_t = 0) we have δϕl(0)ϕl0(0)proportional-to𝛿subscriptitalic-ϕ𝑙0subscriptsuperscriptitalic-ϕ0𝑙0\delta\phi_{l}(0)\propto-\phi^{0}_{l}(0)italic_δ italic_ϕ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ( 0 ) ∝ - italic_ϕ start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ( 0 ) and δϕr(0)ϕr0(0)proportional-to𝛿subscriptitalic-ϕ𝑟0subscriptsuperscriptitalic-ϕ0𝑟0\delta\phi_{r}(0)\propto\phi^{0}_{r}(0)italic_δ italic_ϕ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ( 0 ) ∝ italic_ϕ start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ( 0 ). We find that the gapped Bogoliubov mode corresponds to δnlδnrsimilar-to-or-equals𝛿subscript𝑛𝑙𝛿subscript𝑛𝑟\delta n_{l}\simeq-\delta n_{r}italic_δ italic_n start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ≃ - italic_δ italic_n start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT with δnj=δϕjϕj0𝛿subscript𝑛𝑗𝛿subscriptitalic-ϕ𝑗subscriptsuperscriptitalic-ϕ0𝑗\delta n_{j}=\delta\phi_{j}\phi^{0}_{j}italic_δ italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = italic_δ italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_ϕ start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT, leading to nonzero initial δz𝛿𝑧\delta zitalic_δ italic_z (notice that the initial phase difference δθ=0𝛿𝜃0\delta\theta=0italic_δ italic_θ = 0 since both ϕj0(0)superscriptsubscriptitalic-ϕ𝑗00\phi_{j}^{0}(0)italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT ( 0 ) and δϕj(0)𝛿subscriptitalic-ϕ𝑗0\delta\phi_{j}(0)italic_δ italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( 0 ) are real).

Refer to caption
Figure 7: Wave functions of the ground state and the gapped Bogoliubov excitation. (a) Wave function of ground state ψ0,jssubscript𝜓0𝑗𝑠\psi_{0,js}italic_ψ start_POSTSUBSCRIPT 0 , italic_j italic_s end_POSTSUBSCRIPT (unnormalized with arbitrary unit). (b) Wave function of gapped Bogoliubov excitation δψjs𝛿subscript𝜓𝑗𝑠\delta\psi_{js}italic_δ italic_ψ start_POSTSUBSCRIPT italic_j italic_s end_POSTSUBSCRIPT (unnormalized with arbitrary unit). Parameters are ΩL=0.4ERPlanck-constant-over-2-pisubscriptΩL0.4subscript𝐸R\hbar\Omega_{\text{L}}=0.4\,E_{\text{R}}roman_ℏ roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT = 0.4 italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT and gn=0.25ER𝑔𝑛0.25subscript𝐸Rgn=0.25\,E_{\text{R}}italic_g italic_n = 0.25 italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT.

Explanation of oscillation frequency for ΩL=0Planck-constant-over-2-pisubscriptΩL0\hbar\Omega_{\text{L}}=0roman_ℏ roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT = 0.

The Raman detuning quench technique leads to the excitation of the gapped Bogoliubov mode (corresponding to spin excitation), and the interference with the ground state leads to spin oscillation with an identical frequency as the Bogoliubov gap. In the weak lattice region, the system is well described by the two-mode Josephson model. As one quenches the detuning, the system is quenched away from the fixed point and enters plasma oscillation (i.e., the spin oscillation). Therefore, in the weak lattice region, when ΩLPlanck-constant-over-2-pisubscriptΩL\hbar\Omega_{\text{L}}roman_ℏ roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT is small, both the Bogoliubov gap and the Josephson plasma frequency are determined by the observed spin oscillation. At ΩL=0Planck-constant-over-2-pisubscriptΩL0\hbar\Omega_{\text{L}}=0roman_ℏ roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT = 0, there would be no Josephson oscillation at all, so the numerical two-mode Josephson oscillating frequency is obtained by taking the limit ΩL0Planck-constant-over-2-pisubscriptΩL0\hbar\Omega_{\text{L}}\rightarrow 0roman_ℏ roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT → 0. Experimentally, this gap/frequency is obtained using Bragg spectroscopy instead of quench dynamics, where two Bragg laser beams, collinear with the Raman laser beams of varying detuning, are pulsed onto a SOC BEC. The obtained frequency from this Bragg measurement corresponds to the star marker at ΩL=0Planck-constant-over-2-pisubscriptΩL0\hbar\Omega_{\text{L}}=0roman_ℏ roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT = 0 in Fig. 3 (d) and (e) in the main text. Also, it is worth noting that the real-space Josephson dynamics usually rely on the tight trap and do not have an energy-momentum spectrum. Here, the momentum space Josephson dynamics does not require a real-space trap. Hence, the momentum is a good quantum number, and the pseudo-Goldstone excitation can have an energy-momentum spectrum similar to a pseudo-Goldstone boson.

Additional experimental details.

In our experiments, we have approximately 2.2×1052.2superscript1052.2\times 10^{5}2.2 × 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT atoms of 87Rb confined in a harmonic trap with trap frequencies 𝝎=(ωx,ωy,ωz)=2π(23,154,192)𝝎subscript𝜔𝑥subscript𝜔𝑦subscript𝜔𝑧2𝜋23154192\bm{\omega}=(\omega_{x},\omega_{y},\omega_{z})=2\pi(23,154,192)bold_italic_ω = ( italic_ω start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT , italic_ω start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT , italic_ω start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ) = 2 italic_π ( 23 , 154 , 192 ) Hz. The BEC is prepared in the |1,111\lvert 1,-1\rangle| 1 , - 1 ⟩ Zeeman state. A \qty10G external magnetic field leads to Zeeman splitting of the F=1𝐹1F=1italic_F = 1 states. The quadratic Zeeman shift for the |1,+111\lvert 1,+1\rangle| 1 , + 1 ⟩ state is \qty14.6kHz. Therefore, this state is out of resonance when near-resonant Raman coupling is applied between the |1,111\lvert 1,-1\rangle| 1 , - 1 ⟩ and |1,010\lvert 1,0\rangle| 1 , 0 ⟩ states. The Raman and optical lattice coupling strengths are denoted as ΩRPlanck-constant-over-2-pisubscriptΩR\hbar\Omega_{\text{R}}roman_ℏ roman_Ω start_POSTSUBSCRIPT R end_POSTSUBSCRIPT and ΩLPlanck-constant-over-2-pisubscriptΩL\hbar\Omega_{\text{L}}roman_ℏ roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT respectively, with kRPlanck-constant-over-2-pisubscript𝑘R\hbar k_{\text{R}}roman_ℏ italic_k start_POSTSUBSCRIPT R end_POSTSUBSCRIPT and kLPlanck-constant-over-2-pisubscript𝑘L\hbar k_{\text{L}}roman_ℏ italic_k start_POSTSUBSCRIPT L end_POSTSUBSCRIPT being the corresponding recoil momentum where kL<kRsubscript𝑘Lsubscript𝑘Rk_{\text{L}}<k_{\text{R}}italic_k start_POSTSUBSCRIPT L end_POSTSUBSCRIPT < italic_k start_POSTSUBSCRIPT R end_POSTSUBSCRIPT.

At the end of the experiment, the BEC is released from the crossed optical dipole trap and finally imaged after a 17.5 ms time of flight (TOF). The absorption images resolve the spin-momentum states, as shown in Fig. 2 in the main text. These images involve six atomic clouds, of which two are strongly populated while the others are faint due to the weak optical lattice coupling to these states. The two-photon coupling between the |1,111\lvert 1,-1\rangle| 1 , - 1 ⟩ and |1,010\lvert 1,0\rangle| 1 , 0 ⟩ states effected by the Raman lasers separates the two strongly populated clouds by 2kR2Planck-constant-over-2-pisubscript𝑘R2\hbar k_{\text{R}}2 roman_ℏ italic_k start_POSTSUBSCRIPT R end_POSTSUBSCRIPT in momentum space. Additionally, the optical lattice, coupling the two band minima of the SOC BEC, induces a momentum kick of 2kL2Planck-constant-over-2-pisubscript𝑘L2\hbar k_{\text{L}}2 roman_ℏ italic_k start_POSTSUBSCRIPT L end_POSTSUBSCRIPT in each of the spin states, which effectively leaves the two BECs in the |delimited-|⟩\lvert\uparrow\rangle| ↑ ⟩ and |delimited-|⟩\lvert\downarrow\rangle| ↓ ⟩ spin-states separated by 2(kRkL)2Planck-constant-over-2-pisubscript𝑘Rsubscript𝑘L2\hbar\,(k_{\text{R}}-k_{\text{L}})2 roman_ℏ ( italic_k start_POSTSUBSCRIPT R end_POSTSUBSCRIPT - italic_k start_POSTSUBSCRIPT L end_POSTSUBSCRIPT ) in the undressed picture. The lattice coupling to the other four states in the SOC-dressed and undressed picture is shown via schematic diagrams in [36].

Reasoning for the choice of a finite final detuning.

In our system, the repulsive inter-mode interaction is equivalent to attractive intra-mode interaction. Therefore, a negative curvature is expected at small K𝐾Kitalic_K values if we work at zero detuning, as shown in Fig. 8. When δf=0subscript𝛿𝑓0\delta_{f}=0italic_δ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT = 0, the ground state would be (z0,θ0=0)subscript𝑧0subscript𝜃00\left(z_{0},\,\theta_{0}=0\right)( italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0 ) at large K𝐾Kitalic_K values, where KΩLproportional-to𝐾subscriptΩLK\propto\Omega_{\text{L}}italic_K ∝ roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT. Then, the plasma frequency reads 2K(2KUn)2𝐾2𝐾𝑈𝑛\sqrt{2K\,(2K-Un)}square-root start_ARG 2 italic_K ( 2 italic_K - italic_U italic_n ) end_ARG, which decreases to zero at K=Un/2𝐾𝑈𝑛2K=Un/2italic_K = italic_U italic_n / 2 (i.e., Λ=1Λ1\Lambda=-1roman_Λ = - 1). This indicates a phase transition in the weak lattice regime shown in Fig. 8 by the dashed vertical black line at ΩL0.17ERsimilar-to-or-equalsPlanck-constant-over-2-pisubscriptΩL0.17subscript𝐸𝑅\hbar\Omega_{\text{L}}\simeq 0.17\,E_{R}roman_ℏ roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT ≃ 0.17 italic_E start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT. For K<Un/2𝐾𝑈𝑛2K<Un/2italic_K < italic_U italic_n / 2 (i.e., Λ<1Λ1\Lambda<-1roman_Λ < - 1), the ground state has z02=1Λ20superscriptsubscript𝑧021superscriptΛ20z_{0}^{2}=1-\Lambda^{-2}\neq 0italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 1 - roman_Λ start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT ≠ 0 (z0subscript𝑧0z_{0}italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT can be negative or positive due to the spontaneous symmetry breaking). So, in the weak K𝐾Kitalic_K regime, the plasma frequency reads U2n24K2superscript𝑈2superscript𝑛24superscript𝐾2\sqrt{U^{2}n^{2}-4K^{2}}square-root start_ARG italic_U start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 4 italic_K start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG, which will display a negative curvature and hence a nonlinear effect. In this experiment, we work at a finite final detuning (δf=2π×\qty500Hzsubscript𝛿𝑓2𝜋\qty500𝐻𝑧\delta_{f}=2\pi\times\qty{500}{Hz}italic_δ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT = 2 italic_π × 500 italic_H italic_z) due to the following reasons.

The interaction in our system is not very strong, so its effect is more significant in the weak lattice regime. However, the phase transition occurs at a weak lattice depth of ΩL0.17ERsimilar-to-or-equalsPlanck-constant-over-2-pisubscriptΩL0.17subscript𝐸𝑅\hbar\Omega_{\text{L}}\simeq 0.17\,E_{R}roman_ℏ roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT ≃ 0.17 italic_E start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT. Therefore, in a weak lattice regime, the plasma frequency is very small in a large interval, making it hard to measure with high precision. Also, the ground state preparation is more challenging around the phase-transition point, requiring long ramp times of the parameters. Moreover, considering the \qty100Hz\qty100𝐻𝑧\qty{100}{Hz}100 italic_H italic_z uncertainty in our detuning, the dynamics become more sensitive to the detuning in the weak lattice regime.

Hence, δf=0subscript𝛿𝑓0\delta_{f}=0italic_δ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT = 0 is not a good choice for measuring the oscillation dynamics. Also, choosing a finite final detuning does not prevent us from observing the two types of Josephson oscillation. We chose δf=2π×\qty500Hzsubscript𝛿𝑓2𝜋\qty500𝐻𝑧\delta_{f}=2\pi\times\qty{500}{Hz}italic_δ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT = 2 italic_π × 500 italic_H italic_z, which is not too large but dominates over the uncertainty.

Refer to caption
Figure 8: Comparison of the Bogoliubov spectrum analysis for gn=0.25ER𝑔𝑛0.25subscript𝐸Rgn=0.25\,E_{\text{R}}italic_g italic_n = 0.25 italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT and Raman detuning δ=2π×500𝛿2𝜋500\delta=2\pi\times 500italic_δ = 2 italic_π × 500 Hz (green solid line), 400400400400 Hz (green dashed line), and 300300300300 Hz (green densely dash-dot-dotted line) with the experimental data points for δ=2π×500𝛿2𝜋500\delta=2\pi\times 500italic_δ = 2 italic_π × 500 Hz. The solid orange line represents the variation of ΔΔ\Deltaroman_Δ with respect to ΩLPlanck-constant-over-2-pisubscriptΩL\hbar\Omega_{\text{L}}roman_ℏ roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT for δ=2π×500𝛿2𝜋500\delta=2\pi\times 500italic_δ = 2 italic_π × 500 Hz and gn=0𝑔𝑛0gn=0italic_g italic_n = 0 (i.e., single particle analysis without interactions). The blue solid line is for δ=0𝛿0\delta=0italic_δ = 0 and gn=0.25ER𝑔𝑛0.25subscript𝐸Rgn=0.25\,E_{\text{R}}italic_g italic_n = 0.25 italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT, where a phase transition occurs at ΩL0.17ERsimilar-to-or-equalsPlanck-constant-over-2-pisubscriptΩL0.17subscript𝐸R\hbar\Omega_{\text{L}}\simeq 0.17\,E_{\text{R}}roman_ℏ roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT ≃ 0.17 italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT which is represented by the dashed vertical black line.

Role of atomic interactions.

Refer to caption
Figure 9: (a)–(d) Additional data for the oscillation of σzdelimited-⟨⟩subscript𝜎𝑧\langle\sigma_{z}\rangle⟨ italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ⟩. The solid black curves represent the sinusoidal fitting of the experimental data represented by the distinct markers, and the light blue curves are the GP simulation results. See text for parameters.

The coherent coupling of two macroscopic quantum states gives rise to the Josephson phenomenon. In this experiment, we are working with a residual detuning (ElErsubscript𝐸𝑙subscript𝐸𝑟E_{l}\neq E_{r}italic_E start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ≠ italic_E start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT), and the constant plasma frequency at small K𝐾Kitalic_K (proportional to ΩLsubscriptΩL\Omega_{\text{L}}roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT) values capture the interaction effect for gn=0.25ER𝑔𝑛0.25subscript𝐸Rgn=0.25\,E_{\text{R}}italic_g italic_n = 0.25 italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT. For the non-interacting case (gn=0𝑔𝑛0gn=0italic_g italic_n = 0), shown by the orange line in Fig. 3 (e) in the main text, the plasma frequency would be (ElEr)2+4K2superscriptsubscript𝐸𝑙subscript𝐸𝑟24superscript𝐾2\sqrt{(E_{l}-E_{r})^{2}+4K^{2}}square-root start_ARG ( italic_E start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 4 italic_K start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG in the weak lattice regime. The observed constant plateau cannot be produced from the non-interaction curve by rescaling or shifting experimental parameters.

The real-space atom-atom interaction is repulsive, but in the momentum space, it is effectively attractive. This can be seen by considering the two momentum modes. The inter-mode interaction is repulsive because when both modes are populated, the formed real-space density stripes have high interaction energy, so the BEC prefers to occupy a single momentum mode due to interaction. Effectively, this corresponds to attractive intra-mode interaction (i.e., the effective interaction is negative). As mentioned in the previous section, the interactions modify the dynamics and lead to a constant plateau of plasma frequency in a weak lattice regime, which is observed experimentally. Such a plateau is unique for the negative interaction, making it more difficult to mix the two modes. Thus, the negative interaction will weaken the effects of the coupling K𝐾Kitalic_K, leading to the plateau structure.

Refer to caption
Figure 10: (a) Schematic of the transitions corresponding to the observed resonance peaks in Fig. 11 (a)–(c). (b) A summed image of three absorption images corresponding to the three Bragg resonance peaks in Fig. 11 (a)–(c) showing all the bare spin-momentum eigenstates |idelimited-|⟩𝑖\lvert i\rangle| italic_i ⟩. Box 1, 2, 3, 5, and 6 encloses atoms in |,2kLx^2Planck-constant-over-2-pisubscript𝑘L^𝑥\lvert\uparrow,\,-2\hbar k_{\text{L}}\,\hat{x}\rangle| ↑ , - 2 roman_ℏ italic_k start_POSTSUBSCRIPT L end_POSTSUBSCRIPT over^ start_ARG italic_x end_ARG ⟩, |, 0 0\lvert\uparrow,\,0\rangle| ↑ , 0 ⟩, |, 2kLx^2Planck-constant-over-2-pisubscript𝑘L^𝑥\lvert\uparrow,\,2\hbar k_{\text{L}}\,\hat{x}\rangle| ↑ , 2 roman_ℏ italic_k start_POSTSUBSCRIPT L end_POSTSUBSCRIPT over^ start_ARG italic_x end_ARG ⟩, |,2(kRkL)x^2Planck-constant-over-2-pisubscript𝑘Rsubscript𝑘L^𝑥\lvert\downarrow,\,-2\hbar(k_{\text{R}}-k_{\text{L}})\,\hat{x}\rangle| ↓ , - 2 roman_ℏ ( italic_k start_POSTSUBSCRIPT R end_POSTSUBSCRIPT - italic_k start_POSTSUBSCRIPT L end_POSTSUBSCRIPT ) over^ start_ARG italic_x end_ARG ⟩, and |, 2kRx^2Planck-constant-over-2-pisubscript𝑘R^𝑥\lvert\downarrow,\,2\hbar k_{\text{R}}\,\hat{x}\rangle| ↓ , 2 roman_ℏ italic_k start_POSTSUBSCRIPT R end_POSTSUBSCRIPT over^ start_ARG italic_x end_ARG ⟩ bare state, respectively. The number of atoms in box 4 is negligible.

Experimental parameters.

For our experimental parameters ΩR=2.7ERPlanck-constant-over-2-pisubscriptΩR2.7subscript𝐸R\hbar\Omega_{\text{R}}=2.7\,E_{\text{R}}roman_ℏ roman_Ω start_POSTSUBSCRIPT R end_POSTSUBSCRIPT = 2.7 italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT, ER=1960subscript𝐸R1960E_{\text{R}}=1960italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT = 1960 Hz, δ=2π×\qty500Hz𝛿2𝜋\qty500𝐻𝑧\delta=2\pi\times\qty{500}{Hz}italic_δ = 2 italic_π × 500 italic_H italic_z, and kL=1(ΩR/4ER)2kRsubscript𝑘L1superscriptsubscriptΩR4subscript𝐸R2subscript𝑘Rk_{\text{L}}=\sqrt{1-(\Omega_{\text{R}}/4E_{\text{R}})^{2}}k_{\text{R}}italic_k start_POSTSUBSCRIPT L end_POSTSUBSCRIPT = square-root start_ARG 1 - ( roman_Ω start_POSTSUBSCRIPT R end_POSTSUBSCRIPT / 4 italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG italic_k start_POSTSUBSCRIPT R end_POSTSUBSCRIPT, we find the right band minimum at kmr=0.78kRsubscript𝑘subscript𝑚𝑟0.78subscript𝑘Rk_{m_{r}}=0.78k_{\text{R}}italic_k start_POSTSUBSCRIPT italic_m start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT end_POSTSUBSCRIPT = 0.78 italic_k start_POSTSUBSCRIPT R end_POSTSUBSCRIPT and the left minimum at kml=0.664kRsubscript𝑘subscript𝑚𝑙0.664subscript𝑘Rk_{m_{l}}=-0.664k_{\text{R}}italic_k start_POSTSUBSCRIPT italic_m start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT end_POSTSUBSCRIPT = - 0.664 italic_k start_POSTSUBSCRIPT R end_POSTSUBSCRIPT, so that the ideal lattice has kLideal=0.722kRsuperscriptsubscript𝑘Lideal0.722subscript𝑘Rk_{\text{L}}^{\text{ideal}}=0.722k_{\text{R}}italic_k start_POSTSUBSCRIPT L end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ideal end_POSTSUPERSCRIPT = 0.722 italic_k start_POSTSUBSCRIPT R end_POSTSUBSCRIPT. However, in the realized experiment, we have used kL=1(ΩR/4ER)2kR=0.7378kRsubscript𝑘L1superscriptsubscriptΩR4subscript𝐸R2subscript𝑘R0.7378subscript𝑘Rk_{\text{L}}=\sqrt{1-(\Omega_{\text{R}}/4E_{\text{R}})^{2}}k_{\text{R}}=0.7378% k_{\text{R}}italic_k start_POSTSUBSCRIPT L end_POSTSUBSCRIPT = square-root start_ARG 1 - ( roman_Ω start_POSTSUBSCRIPT R end_POSTSUBSCRIPT / 4 italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG italic_k start_POSTSUBSCRIPT R end_POSTSUBSCRIPT = 0.7378 italic_k start_POSTSUBSCRIPT R end_POSTSUBSCRIPT. Since the BEC is initially prepared with momentum kmrsubscript𝑘subscript𝑚𝑟k_{m_{r}}italic_k start_POSTSUBSCRIPT italic_m start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT end_POSTSUBSCRIPT, the two modes should be at momenta kmr=0.78kRsubscript𝑘subscript𝑚𝑟0.78subscript𝑘Rk_{m_{r}}=0.78k_{\text{R}}italic_k start_POSTSUBSCRIPT italic_m start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT end_POSTSUBSCRIPT = 0.78 italic_k start_POSTSUBSCRIPT R end_POSTSUBSCRIPT and kmr2kL=0.6956kRsubscript𝑘subscript𝑚𝑟2subscript𝑘L0.6956subscript𝑘Rk_{m_{r}}-2k_{\text{L}}=-0.6956k_{\text{R}}italic_k start_POSTSUBSCRIPT italic_m start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT end_POSTSUBSCRIPT - 2 italic_k start_POSTSUBSCRIPT L end_POSTSUBSCRIPT = - 0.6956 italic_k start_POSTSUBSCRIPT R end_POSTSUBSCRIPT. Then we have χr=[0.331,0.9436]Tsubscript𝜒𝑟superscript0.3310.9436𝑇\chi_{r}=[0.331,-0.9436]^{T}italic_χ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT = [ 0.331 , - 0.9436 ] start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT and χl=[0.9123,0.4096]Tsubscript𝜒𝑙superscript0.91230.4096𝑇\chi_{l}=[0.9123,-0.4096]^{T}italic_χ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT = [ 0.9123 , - 0.4096 ] start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT. ΔE=0.1874ER/(2K)Δ𝐸0.1874subscript𝐸R2𝐾\Delta E=-0.1874E_{\text{R}}/(2K)roman_Δ italic_E = - 0.1874 italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT / ( 2 italic_K ), Λ=0.4613gn/(2K)Λ0.4613𝑔𝑛2𝐾\Lambda=-0.4613gn/(2K)roman_Λ = - 0.4613 italic_g italic_n / ( 2 italic_K ), 2K=0.6792ΩL2𝐾0.6792subscriptΩL2K=0.6792\Omega_{\text{L}}2 italic_K = 0.6792 roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT. For a given gn𝑔𝑛gnitalic_g italic_n and ΩLsubscriptΩL\Omega_{\text{L}}roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT, we can solve for z0subscript𝑧0z_{0}italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and obtain the oscillation frequency. The corresponding polarization is σz=a+bzdelimited-⟨⟩subscript𝜎𝑧𝑎𝑏𝑧\langle\sigma_{z}\rangle=a+bz⟨ italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ⟩ = italic_a + italic_b italic_z with coefficients a=0.0493,b=0.732formulae-sequence𝑎0.0493𝑏0.732a=0.0493,b=0.732italic_a = 0.0493 , italic_b = 0.732.

We note that the dynamics are not sensitive to kmrsubscript𝑘subscript𝑚𝑟k_{m_{r}}italic_k start_POSTSUBSCRIPT italic_m start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT end_POSTSUBSCRIPT and kLsubscript𝑘Lk_{\text{L}}italic_k start_POSTSUBSCRIPT L end_POSTSUBSCRIPT, so the effect of slight deviations of kLsubscript𝑘Lk_{\text{L}}italic_k start_POSTSUBSCRIPT L end_POSTSUBSCRIPT from the ideal value is negligible. On the other hand, the ideal kLsubscript𝑘Lk_{\text{L}}italic_k start_POSTSUBSCRIPT L end_POSTSUBSCRIPT depends on the Raman detuning. Here, we set kLsubscript𝑘Lk_{\text{L}}italic_k start_POSTSUBSCRIPT L end_POSTSUBSCRIPT to be the ideal value corresponding to zero detuning.

Additional plasma oscillation results.

In Fig. 9, we present additional results corresponding to the Bogoliubov spectrum using the Raman detuning quench technique. This demonstrates the robust nature of our experiment, which provides access to study the dependence of ΔΔ\Deltaroman_Δ on ΩLPlanck-constant-over-2-pisubscriptΩL\hbar\Omega_{\text{L}}roman_ℏ roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT for a range of ΩLPlanck-constant-over-2-pisubscriptΩL\hbar\Omega_{\text{L}}roman_ℏ roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT. The corresponding Q,ΩL,Δ𝑄Planck-constant-over-2-pisubscriptΩLΔQ,\,\hbar\Omega_{\text{L}},\,\Deltaitalic_Q , roman_ℏ roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT , roman_Δ (see main text for definition) are (a) (2π×\qty1.0kHz,0.4ER,(0.287±0.008)ER)2𝜋\qty1.0𝑘𝐻𝑧0.4subscript𝐸Rplus-or-minus0.2870.008subscript𝐸R(2\pi\times\qty{1.0}{kHz},0.4\,E_{\text{R}},(0.287\pm 0.008)\,E_{\text{R}})( 2 italic_π × 1.0 italic_k italic_H italic_z , 0.4 italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT , ( 0.287 ± 0.008 ) italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT ), (b) (2π×\qty900Hz,0.6ER,(0.382±0.004)ER)2𝜋\qty900𝐻𝑧0.6subscript𝐸Rplus-or-minus0.3820.004subscript𝐸R(2\pi\times\qty{900}{Hz},0.6\,E_{\text{R}},(0.382\pm 0.004)\,E_{\text{R}})( 2 italic_π × 900 italic_H italic_z , 0.6 italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT , ( 0.382 ± 0.004 ) italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT ), (c) (2π×\qty500Hz,0.8ER,(0.438±0.012)ER)2𝜋\qty500𝐻𝑧0.8subscript𝐸Rplus-or-minus0.4380.012subscript𝐸R(2\pi\times\qty{500}{Hz},0.8\,E_{\text{R}},(0.438\pm 0.012)\,E_{\text{R}})( 2 italic_π × 500 italic_H italic_z , 0.8 italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT , ( 0.438 ± 0.012 ) italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT ), and (d) (2π×\qty500Hz,1.2ER,(0.628±0.02)ER)2𝜋\qty500𝐻𝑧1.2subscript𝐸Rplus-or-minus0.6280.02subscript𝐸R(2\pi\times\qty{500}{Hz},1.2\,E_{\text{R}},(0.628\pm 0.02)\,E_{\text{R}})( 2 italic_π × 500 italic_H italic_z , 1.2 italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT , ( 0.628 ± 0.02 ) italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT ). Each experimental data point presented here and in the main text is an average over three measurements, and the error bars represent the standard deviation over those measurements.

Here, we are primarily interested in short-time Josephson oscillations. Within this small time scale (similar-to\sim 2 to 4 ms), decoherence induced by imperfections such as the atom loss and magnetic field fluctuation still have small effects, and coherent oscillation of the spin-polarization (σzdelimited-⟨⟩subscript𝜎𝑧\langle\sigma_{z}\rangle⟨ italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ⟩) is observed. At long times, the magnetic fluctuations and atom-atom scatterings would break the coherence between BECs at the two band minima, leading to damped spin oscillation for long-time dynamics.

Refer to caption
Figure 11: A plot of Nfracdelimited-⟨⟩superscript𝑁frac\langle N^{\text{frac}}\rangle⟨ italic_N start_POSTSUPERSCRIPT frac end_POSTSUPERSCRIPT ⟩ with respect to ΔνBraggΔsubscript𝜈Bragg\Delta\nu_{\text{Bragg}}roman_Δ italic_ν start_POSTSUBSCRIPT Bragg end_POSTSUBSCRIPT for a SOC BEC under Bragg spectroscopy. The parameters are ΩR=2.7ERPlanck-constant-over-2-pisubscriptΩR2.7subscript𝐸R\hbar\Omega_{\text{R}}=2.7\,E_{\text{R}}roman_ℏ roman_Ω start_POSTSUBSCRIPT R end_POSTSUBSCRIPT = 2.7 italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT and δ=2π×\qty500Hz𝛿2𝜋\qty500𝐻𝑧\delta=2\pi\times\qty{500}{Hz}italic_δ = 2 italic_π × 500 italic_H italic_z. Panels (a)-(f) correspond to box no. 1, 3, 5, 2, 4, and 6 in Fig. 10 (b), and the red lines are Gaussian fits to the peaks T3, T2, T1, and T4. Here, each gray circle is an average of over three measurements. An error bar represents the standard deviation of the three measurements. The observed resonance frequencies corresponding to the peaks are (7.432±0.062)plus-or-minus7.4320.062(-7.432\pm 0.062)( - 7.432 ± 0.062 ) kHz, (4.193±0.031)plus-or-minus4.1930.031(4.193\pm 0.031)( 4.193 ± 0.031 ) kHz, (0.495±0.044)plus-or-minus0.4950.044(-0.495\pm 0.044)( - 0.495 ± 0.044 ) kHz, and (4.448±0.141)plus-or-minus4.4480.141(4.448\pm 0.141)( 4.448 ± 0.141 ) kHz in (a)-(c), respectively. Panel (d) corresponds to Box 2, where most atoms were before pulsing the optical lattice. After the optical lattice pulse, atoms transfer to the other states enclosed by other boxes, and hence, there is a dip in Nfracdelimited-⟨⟩superscript𝑁frac\langle N^{\text{frac}}\rangle⟨ italic_N start_POSTSUPERSCRIPT frac end_POSTSUPERSCRIPT ⟩. Panels (e) and (f) show that the atom numbers are negligible in Box 4 and 6.

Bragg spectroscopy.

Experimentally, ΩL=0.2ERPlanck-constant-over-2-pisubscriptΩL0.2subscript𝐸R\hbar\Omega_{\text{L}}=0.2\,E_{\text{R}}roman_ℏ roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT = 0.2 italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT is the lowest lattice strength for which a clearly measurable quench response could be obtained. To extend our measurements to zero lattice strength, Bragg spectroscopy is employed [49]. For this, an optical lattice of small coupling strength (ΩL=0.13ERPlanck-constant-over-2-pisubscriptΩL0.13subscript𝐸R\hbar\Omega_{\text{L}}=0.13\,E_{\text{R}}roman_ℏ roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT = 0.13 italic_E start_POSTSUBSCRIPT R end_POSTSUBSCRIPT) and of a varying detuning (ΔνBraggΔsubscript𝜈Bragg\Delta\nu_{\text{Bragg}}roman_Δ italic_ν start_POSTSUBSCRIPT Bragg end_POSTSUBSCRIPT) is pulsed onto the SOC BEC for 1111 ms.

The fractional population Nifracsubscriptsuperscript𝑁frac𝑖N^{\text{frac}}_{i}italic_N start_POSTSUPERSCRIPT frac end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT of each bare spin-momentum eigenstate |idelimited-|⟩𝑖\lvert i\rangle| italic_i ⟩ is measured after \qty17.5ms of TOF. It is defined as Nifrac=Ni/Ntotalsubscriptsuperscript𝑁frac𝑖subscript𝑁𝑖subscript𝑁totalN^{\text{frac}}_{i}=N_{i}/N_{\text{total}}italic_N start_POSTSUPERSCRIPT frac end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_N start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT / italic_N start_POSTSUBSCRIPT total end_POSTSUBSCRIPT, where Nisubscript𝑁𝑖N_{i}italic_N start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is the number of atoms in the eigenstate |iket𝑖|i\rangle| italic_i ⟩ and Ntotalsubscript𝑁totalN_{\text{total}}italic_N start_POSTSUBSCRIPT total end_POSTSUBSCRIPT is the total number of atoms (Fig. 10 (b)). Three experimental data sets are collected, and the expectation value of the fractional population Nifracdelimited-⟨⟩subscriptsuperscript𝑁frac𝑖\langle N^{\text{frac}}_{i}\rangle⟨ italic_N start_POSTSUPERSCRIPT frac end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⟩ is computed for each ΔνBraggΔsubscript𝜈Bragg\Delta\nu_{\text{Bragg}}roman_Δ italic_ν start_POSTSUBSCRIPT Bragg end_POSTSUBSCRIPT. In Fig. 11, a plot of Nifracdelimited-⟨⟩subscriptsuperscript𝑁frac𝑖\langle N^{\text{frac}}_{i}\rangle⟨ italic_N start_POSTSUPERSCRIPT frac end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⟩ with respect to ΔνBraggΔsubscript𝜈Bragg\Delta\nu_{\text{Bragg}}roman_Δ italic_ν start_POSTSUBSCRIPT Bragg end_POSTSUBSCRIPT is shown, where each observed peak can be explained by referring to the transitions shown in Fig. 10 (a) using the single-particle SOC band structure. The observed resonant frequencies \qty-0.495kHz, \qty4.193kHz, \qty-7.432kHz, and \qty4.448kHz correspond to the transitions T1, T2, T3, and T4 in Fig. 10 (a), where the (±)plus-or-minus(\pm)( ± ) signs refers to positive and negative ΔνBraggΔsubscript𝜈Bragg\Delta\nu_{\text{Bragg}}roman_Δ italic_ν start_POSTSUBSCRIPT Bragg end_POSTSUBSCRIPT. At resonance, a positive detuning imparts a 2kLx^2Planck-constant-over-2-pisubscript𝑘L^𝑥2\hbar k_{\text{L}}\,\hat{x}2 roman_ℏ italic_k start_POSTSUBSCRIPT L end_POSTSUBSCRIPT over^ start_ARG italic_x end_ARG momentum kick, whereas a negative detuning imparts a 2kLx^2Planck-constant-over-2-pisubscript𝑘L^𝑥-2\hbar k_{\text{L}}\,\hat{x}- 2 roman_ℏ italic_k start_POSTSUBSCRIPT L end_POSTSUBSCRIPT over^ start_ARG italic_x end_ARG momentum kick. According to the single-particle SOC band structure, the predicted resonant frequencies for the transitions T1, T2, T3, and T4 are \qty-0.368kHz, \qty3.61kHz, \qty-7.645kHz, and \qty4.448kHz respectively. The difference between the theoretical values and experimental observations is attributed to the mean-field interactions. The measured energy h×|(0.495±0.044)|plus-or-minus0.4950.044h\times\lvert(-0.495\pm 0.044)\rvertitalic_h × | ( - 0.495 ± 0.044 ) | kHz associated with the transition T1 is the limiting value of the Josephson plasma frequency and the zero quasimomentum band gap for ΩL=0Planck-constant-over-2-pisubscriptΩL0\hbar\Omega_{\text{L}}=0roman_ℏ roman_Ω start_POSTSUBSCRIPT L end_POSTSUBSCRIPT = 0.

References

  • Josephson [1962] B. Josephson, Possible new effects in superconductive tunnelling, Physics Letters 1, 251 (1962).
  • Josephson [1974] B. D. Josephson, The discovery of tunnelling supercurrents, Rev. Mod. Phys. 46, 251 (1974).
  • Anderson and Rowell [1963] P. W. Anderson and J. M. Rowell, Probable observation of the Josephson superconducting tunneling effect, Phys. Rev. Lett. 10, 230 (1963).
  • Backhaus et al. [1997] S. Backhaus, S. V. Pereverzev, A. Loshak, J. C. Davis, and R. E. Packard, Direct measurement of the current-phase relation of a superfluid 3He-B weak link, Science 278, 1435 (1997).
  • Wheatley [1975] J. C. Wheatley, Experimental properties of superfluid He3superscriptHe3{}^{3}\mathrm{He}start_FLOATSUPERSCRIPT 3 end_FLOATSUPERSCRIPT roman_HeRev. Mod. Phys. 47, 415 (1975).
  • Leggett [1975] A. J. Leggett, A theoretical description of the new phases of liquid He3superscriptHe3{}^{3}\mathrm{He}start_FLOATSUPERSCRIPT 3 end_FLOATSUPERSCRIPT roman_HeRev. Mod. Phys. 47, 331 (1975).
  • Sukhatme et al. [2001] K. Sukhatme, Y. Mukharsky, T. Chui, and D. Pearson, Observation of the ideal Josephson effect in superfluid He4superscriptHe4{}^{4}\mathrm{He}start_FLOATSUPERSCRIPT 4 end_FLOATSUPERSCRIPT roman_HeNature 411, 280 (2001).
  • Hoskinson et al. [2005] E. Hoskinson, R. Packard, and T. M. Haard, Quantum whistling in superfluid helium-4, Nature 433, 376 (2005).
  • Abbarchi et al. [2013] M. Abbarchi, A. Amo, V. Sala, D. Solnyshkov, H. Flayac, L. Ferrier, I. Sagnes, E. Galopin, A. Lemaître, G. Malpuech, et al., Macroscopic quantum self-trap** and Josephson oscillations of exciton polaritons, Nature Physics 9, 275 (2013).
  • Albiez et al. [2005] M. Albiez, R. Gati, J. Fölling, S. Hunsmann, M. Cristiani, and M. K. Oberthaler, Direct observation of tunneling and nonlinear self-trap** in a single bosonic Josephson junction, Phys. Rev. Lett. 95, 010402 (2005).
  • Levy et al. [2007] S. Levy, E. Lahoud, I. Shomroni, and J. Steinhauer, The ac and dc Josephson effects in a Bose-Einstein condensate, Nature 449, 579 (2007).
  • Dalfovo et al. [1996] F. Dalfovo, L. Pitaevskii, and S. Stringari, Order parameter at the boundary of a trapped Bose gas, Phys. Rev. A 54, 4213 (1996).
  • Dalfovo et al. [1999] F. Dalfovo, S. Giorgini, L. P. Pitaevskii, and S. Stringari, Theory of Bose-Einstein condensation in trapped gases, Rev. Mod. Phys. 71, 463 (1999).
  • Andrews et al. [1997] M. R. Andrews, C. G. Townsend, H.-J. Miesner, D. S. Durfee, D. M. Kurn, and W. Ketterle, Observation of interference between two Bose condensates, Science 275, 637 (1997).
  • Smerzi et al. [1997] A. Smerzi, S. Fantoni, S. Giovanazzi, and S. R. Shenoy, Quantum coherent atomic tunneling between two trapped Bose-Einstein condensates, Phys. Rev. Lett. 79, 4950 (1997).
  • Öhberg and Stenholm [1999] P. Öhberg and S. Stenholm, Internal Josephson effect in trapped double condensates, Phys. Rev. A 59, 3890 (1999).
  • Williams et al. [1999] J. Williams, R. Walser, J. Cooper, E. Cornell, and M. Holland, Nonlinear Josephson-type oscillations of a driven, two-component Bose-Einstein condensate, Phys. Rev. A 59, R31 (1999).
  • Raghavan et al. [1999] S. Raghavan, A. Smerzi, S. Fantoni, and S. R. Shenoy, Coherent oscillations between two weakly coupled Bose-Einstein condensates: Josephson effects, π𝜋\piitalic_π oscillations, and macroscopic quantum self-trap**, Phys. Rev. A 59, 620 (1999).
  • Cataliotti et al. [2001] F. S. Cataliotti, S. Burger, C. Fort, P. Maddaloni, F. Minardi, A. Trombettoni, A. Smerzi, and M. Inguscio, Josephson junction arrays with Bose-Einstein condensates, Science 293, 843 (2001).
  • Zibold et al. [2010] T. Zibold, E. Nicklas, C. Gross, and M. K. Oberthaler, Classical bifurcation at the transition from Rabi to Josephson dynamics, Phys. Rev. Lett. 105, 204101 (2010).
  • Kreula et al. [2017] J. M. Kreula, G. Valtolina, and P. Törmä, Spin-asymmetric Josephson plasma oscillations, Phys. Rev. A 95, 013634 (2017).
  • Spagnolli et al. [2017] G. Spagnolli, G. Semeghini, L. Masi, G. Ferioli, A. Trenkwalder, S. Coop, M. Landini, L. Pezzè, G. Modugno, M. Inguscio, A. Smerzi, and M. Fattori, Crossing over from attractive to repulsive interactions in a tunneling bosonic Josephson junction, Phys. Rev. Lett. 118, 230403 (2017).
  • Burchianti et al. [2017] A. Burchianti, C. Fort, and M. Modugno, Josephson plasma oscillations and the Gross-Pitaevskii equation: Bogoliubov approach versus two-mode model, Phys. Rev. A 95, 023627 (2017).
  • Burchianti et al. [2018] A. Burchianti, F. Scazza, A. Amico, G. Valtolina, J. A. Seman, C. Fort, M. Zaccanti, M. Inguscio, and G. Roati, Connecting dissipation and phase slips in a Josephson junction between fermionic superfluids, Phys. Rev. Lett. 120, 025302 (2018).
  • Valtolina et al. [2015] G. Valtolina, A. Burchianti, A. Amico, E. Neri, K. Xhani, J. A. Seman, A. Trombettoni, A. Smerzi, M. Zaccanti, M. Inguscio, and G. Roati, Josephson effect in fermionic superfluids across the BEC-BCS crossover, Science 350, 1505 (2015).
  • Luick et al. [2020] N. Luick, L. Sobirey, M. Bohlen, V. P. Singh, L. Mathey, T. Lompe, and H. Moritz, An ideal Josephson junction in an ultracold two-dimensional Fermi gas, Science 369, 89 (2020).
  • Makhlin et al. [2001] Y. Makhlin, G. Schön, and A. Shnirman, Quantum-state engineering with Josephson-junction devices, Rev. Mod. Phys. 73, 357 (2001).
  • Ryu et al. [2013] C. Ryu, P. W. Blackburn, A. A. Blinova, and M. G. Boshier, Experimental realization of Josephson junctions for an atom SQUID, Phys. Rev. Lett. 111, 205301 (2013).
  • Martinis et al. [2002] J. M. Martinis, S. Nam, J. Aumentado, and C. Urbina, Rabi oscillations in a large Josephson-junction qubit, Phys. Rev. Lett. 89, 117901 (2002).
  • Astafiev et al. [2006] O. Astafiev, Y. A. Pashkin, Y. Nakamura, T. Yamamoto, and J. S. Tsai, Temperature square dependence of the low frequency 1/f1𝑓1/f1 / italic_f charge noise in the Josephson junction qubits, Phys. Rev. Lett. 96, 137001 (2006).
  • Martinis et al. [2009] J. M. Martinis, M. Ansmann, and J. Aumentado, Energy decay in superconducting Josephson-junction qubits from nonequilibrium quasiparticle excitations, Phys. Rev. Lett. 103, 097002 (2009).
  • Paik et al. [2011] H. Paik, D. I. Schuster, L. S. Bishop, G. Kirchmair, G. Catelani, A. P. Sears, B. R. Johnson, M. J. Reagor, L. Frunzio, L. I. Glazman, S. M. Girvin, M. H. Devoret, and R. J. Schoelkopf, Observation of high coherence in Josephson junction qubits measured in a three-dimensional circuit QED architecture, Phys. Rev. Lett. 107, 240501 (2011).
  • Hou et al. [2018] J. Hou, X.-W. Luo, K. Sun, T. Bersano, V. Gokhroo, S. Mossman, P. Engels, and C. Zhang, Momentum-space Josephson effects, Phys. Rev. Lett. 120, 120401 (2018).
  • An et al. [2018] F. A. An, E. J. Meier, J. Ang’ong’a, and B. Gadway, Correlated dynamics in a synthetic lattice of momentum states, Phys. Rev. Lett. 120, 040407 (2018).
  • An et al. [2021] F. A. An, B. Sundar, J. Hou, X.-W. Luo, E. J. Meier, C. Zhang, K. R. A. Hazzard, and B. Gadway, Nonlinear dynamics in a synthetic momentum-state lattice, Phys. Rev. Lett. 127, 130401 (2021).
  • Bersano et al. [2019] T. M. Bersano, J. Hou, S. Mossman, V. Gokhroo, X.-W. Luo, K. Sun, C. Zhang, and P. Engels, Experimental realization of a long-lived striped Bose-Einstein condensate induced by momentum-space hop**, Phys. Rev. A 99, 051602 (2019).
  • Lin et al. [2011] Y.-J. Lin, K. Jiménez-García, and I. B. Spielman, Spin-orbit-coupled Bose-Einstein condensates, Nature 471, 83 (2011).
  • Jiménez-García et al. [2015] K. Jiménez-García, L. J. LeBlanc, R. A. Williams, M. C. Beeler, C. Qu, M. Gong, C. Zhang, and I. B. Spielman, Tunable spin-orbit coupling via strong driving in ultracold-atom systems, Phys. Rev. Lett. 114, 125301 (2015).
  • Zhang et al. [2016] Y. Zhang, M. E. Mossman, T. Busch, P. Engels, and C. Zhang, Properties of spin-orbit-coupled Bose-Einstein condensates, Frontiers of Physics 11, 118103 (2016).
  • Burgess [2000] C. Burgess, Goldstone and pseudo-goldstone bosons in nuclear, particle and condensed-matter physics, Physics Reports 330, 193 (2000).
  • Weinberg [1972] S. Weinberg, Approximate symmetries and pseudo-goldstone bosons, Phys. Rev. Lett. 29, 1698 (1972).
  • [42] .
  • Li et al. [2013] Y. Li, G. I. Martone, L. P. Pitaevskii, and S. Stringari, Superstripes and the excitation spectrum of a spin-orbit-coupled Bose-Einstein condensate, Phys. Rev. Lett. 110, 235302 (2013).
  • Li et al. [2021] G.-Q. Li, X.-W. Luo, J. Hou, and C. Zhang, Pseudo-Goldstone excitations in a striped Bose-Einstein condensate, Phys. Rev. A 104, 023311 (2021).
  • Esposito et al. [2007] F. P. Esposito, L.-P. Guay, R. B. MacKenzie, M. B. Paranjape, and L. C. R. Wijewardhana, Field theoretic description of the abelian and non-abelian Josephson effect, Phys. Rev. Lett. 98, 241602 (2007).
  • Wang et al. [2010] C. Wang, C. Gao, C.-M. Jian, and H. Zhai, Spin-orbit coupled spinor Bose-Einstein condensates, Phys. Rev. Lett. 105, 160403 (2010).
  • Ho and Zhang [2011] T.-L. Ho and S. Zhang, Bose-Einstein condensates with spin-orbit interaction, Phys. Rev. Lett. 107, 150403 (2011).
  • Li et al. [2012] Y. Li, L. P. Pitaevskii, and S. Stringari, Quantum tricriticality and phase transitions in spin-orbit coupled Bose-Einstein condensates, Phys. Rev. Lett. 108, 225301 (2012).
  • Khamehchi et al. [2014] M. A. Khamehchi, Y. Zhang, C. Hamner, T. Busch, and P. Engels, Measurement of collective excitations in a spin-orbit-coupled Bose-Einstein condensate, Phys. Rev. A 90, 063624 (2014).